Next Article in Journal
Applicability of Honey on Silkworms (Bombyx mori) and Quality Improvement of Its Biomaterials
Next Article in Special Issue
Measurement of Adult Human Brain Responses to Breath-Holding by Multi-Distance Hyperspectral Near-Infrared Spectroscopy
Previous Article in Journal
Energy Management and Control System Design of an Integrated Flywheel Energy Storage System for Residential Users
Previous Article in Special Issue
Dual-Slope Diffuse Reflectance Instrument for Calibration-Free Broadband Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Use of Supercontinuum Laser Sources in Biomedical Diffuse Optics: Unlocking the Power of Multispectral Imaging

1
Biomedical Optics Research Laboratory, Department of Medical Physics and Biomedical Engineering, University College London, London WC1E 6BT, UK
2
National Institute of Optics, National Research Council, 50019 Sesto Fiorentino, Italy
3
European Laboratory for Non-linear Spectroscopy, 50019 Sesto Fiorentino, Italy
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(10), 4616; https://doi.org/10.3390/app11104616
Submission received: 11 April 2021 / Revised: 5 May 2021 / Accepted: 7 May 2021 / Published: 18 May 2021

Abstract

:
Optical techniques based on diffuse optics have been around for decades now and are making their way into the day-to-day medical applications. Even though the physics foundations of these techniques have been known for many years, practical implementation of these technique were hindered by technological limitations, mainly from the light sources and/or detection electronics. In the past 20 years, the developments of supercontinuum laser (SCL) enabled to unlock some of these limitations, enabling the development of system and methodologies relevant for medical use, notably in terms of spectral monitoring. In this review, we focus on the use of SCL in biomedical diffuse optics, from instrumentation and methods developments to their use for medical applications. A total of 95 publications were identified, from 1993 to 2021. We discuss the advantages of the SCL to cover a large spectral bandwidth with a high spectral power and fast switching against the disadvantages of cost, bulkiness, and long warm up times. Finally, we summarize the utility of using such light sources in the development and application of diffuse optics in biomedical sciences and clinical applications.

1. Introduction

Non-invasive optical monitoring of the human physiology has been a great asset in the medical diagnosis over the last decades [1], and this field is still expanding as technological advances drives the development of new instruments for use in various medical applications [2]. One of the main techniques of optical monitoring of biological tissues is near-infrared spectroscopy (NIRS). It is based on the use of light between 650 and 900 nm, which is within the so-called “transparency window”. Within this light window, the optical absorption of biological tissue is at its lowest and light scattering is sufficiently high so the propagation of light in tissue is facilitated. Therefore, NIRS provides a way to gather information from deep into the tissue [3]. As one of the main absorbers in this window is the hemoglobin, in its oxygenated and deoxygenated forms (oxy- and deoxy-hemoglobin ([HbO2] and [HHb], respectively), NIRS has been broadly used to quantify the changes in oxygenation of various tissues like muscle and brain [4]. This ability to quantify oxygenation non-invasively, and in real time has helped NIRS to become a standard technique in the labs. However, a few pitfalls are still holding back its wide adoption at a clinical level. Indeed, most of the widespread techniques used for NIRS are based on the so-called continuous wave (CW) technique where the tissue is illuminated with a non-coherent light and the variation in light intensity is collected at the detector [5]. In this configuration, it is harder to estimate absolute quantities, as the underlying optical properties of the tissue (mainly absorption and scattering coefficient) cannot be measured, and the scattering properties are assumed to be fixed (i.e., in time and for every subject). Therefore, most of the CW-NIRS systems only quantify the changes in hemoglobin concentration. This can be a drawback for clinical use, as an absolute comparison between two patients cannot be done unless you involve a well-controlled physiological stimulus. The other drawback is that the measurement is often performed with only two wavelengths, the minimum necessary to retrieve two absorbing compounds [HbO2] and [HHb]. However, the contribution of other chromophores is either neglected (like collagen), or assumed and fixed, like water. This results in an increase in the physiological noise, which reduces the accuracy of the measurement.
These drawbacks have been tackled over the recent years, and some new techniques and instrumentation enabled researchers to work around these drawbacks. Firstly, one can use more advanced CW-NIRS techniques in order to extract absolute information about the tissue. We can cite the broadband NIRS that also has the capability to extract absolute information of tissue optical properties [6]. The added benefits of this technique are that, as it uses a large number of wavelength (typically more than a hundred), it can perform a true spectroscopic analysis, hence also gathering information about other chromophores like cytochromes, water, and fat [6]. This extra information enables having a complete optical view of the tissue under investigations. Moreover, on top of providing a more precise view of the oxygenation, the use of many wavelengths enables targeting novel optical contrasts, like the oxidative state of the cytochrome-c-oxidase (oxCCO) which is a biomarker of metabolism [7].
However, even though these improvements are promising, performing a true spectroscopic measurement on tissue using NIRS is still challenging. One of the main issues is the penetration depth of the light and the contamination of the tissues of interest by the upper layers (i.e., skin) [3]. Indeed, the tissues are often modelled as homogeneous for simplicity, even though this hypothesis is not true. Therefore, new methodologies are required to enhance the discrimination between different tissue types, thus improving the accuracy of the measurement.
It is well known that the time-domain NIRS (TD-NIRS) is the most advanced and accurate type of measurement, capable of retrieving absolute information and to discriminate between shallow and deep tissue [8]. Briefly, in TD-NIRS an ultra-short light impulse (typically in the picosecond range) illuminates the tissue. Tissue has the effect of attenuating and broadening the re-emitted pulse (the resulting pulse-width is typically few nanoseconds wide). The measured signal is typically called a temporal point spread function (TPSF) or distribution of time of flights (DTOFs) of photons. Indeed, what is measured here is the histogram of the arrival time of photons. Thus, the amount of information acquired by TD-NIRS systems outstrips the classical CW-NIRS, which only acquires the intensity. From these DTOF, various data processing methods can be employed to extract information about the tissue. It is out of the scope of this paper to detail these methods, but the interested reader can refer to the following reviews for a more in-depth description of TD-NIRS methodologies [9,10], and for a special focus on broadband TD-NIRS [11], which will be largely discussed in the present work.
The extraction of more accurate information with TD-NIRS comes at the cost of a more complex instrumentation, requiring pulsed-laser source (pulse in order of MHz) and fast electronic and sensitive detectors in order to be able to unlock single photon counting. Thus, even though TD-NIRS has been around for many decades now and has been developed in parallel to CW-NIRS, these more specific requirements have hindered its development. Nevertheless, much progress has been made over the last two decades, and TD-NIRS is closing the gap with CW-NIRS in terms of user’s adoption.
One of the key developments that has promoted the development of TD-NIRS is the supercontinuum laser (SCL) sources, which provide a white laser source, with a high power (several watts) and a very short pulse width (typical less than 10 ps) at several tens of MHz. Indeed, these characteristics are essential in order to develop reliable TD-NIRS systems, that started to be developed in the nineties. During these early days of development, no commercial pulsed laser diodes were available, and one had to rely on homemade SC generation or Ti-Sapphire lasers in order to be able to produce a coherence pulsed light with a spectral power density sufficient to be detected by the photon counting electronic. Therefore, SCL sources have been a true driver of TD-NIRS development in the early days, and have more recently unlocked new horizons like spectroscopy, which have broadened the fields of applications.
Finally, besides traditional sparse measurement of tissue properties with diffuse optics, Diffuse Optical Tomography (DOT) has also been rapidly expending in the past decades. This imaging technique is based on the acquisition of a large number of channels, coupled to a more refined data analysis, based on anatomically realistic models of tissue. This allows to reconstruct 3D spatial distribution of optical/physiological parameters of the tissues, rather than providing tissue information on a limited number of points. An extension of this technique is called fluorescence DOT, that uses extrinsic near-infrared fluorescent dyes which targets specific tissues to harvest extra information and enable molecular imaging. The interested reader can refer to reviews [12,13] for more information about these techniques. We will see that SCL sources have also a role to play in order to improve the performances of the DOT instruments.
In this work, we will review the use of SCL in the field of biomedical diffuse optics. Thus, we will begin by a short overview of the history and basic principle of supercontinuum generation. Then, we will review works that have been using SCL in biomedical diffuse optics, from the instrument development to the application. Finally, we will summarize and highlight the strength and weakness of SCL for these applications and show what areas can benefits from its use.

2. History and Basic Principles of Supercontinuum Lasers

Supercontinuum (SC) lasers are broadband, coherent light sources that are capable of combining the wide spectral range of conventional white lamps (including UV, visible and IR light) with the optical performances of single-mode lasers, in particular their high output intensity, the capability of fast, pulsed emission, and the fine collimation of the beam down to diffraction limit. Such characteristics make this type of photonic technology highly versatile for a number of applications, in particular in biomedical optics and biophotonics, thanks also to recent advancement towards making SC lasers more compact and affordable [14].
The physical process of optical SC generation was discovered and studied in the 60 s and 70 s with milestones publications by Alfano and Shapiro [15,16]. It is based on the propagation of extreme non-linear effects from a narrow-bandwidth pulsed laser beam in order to generate a broadening of the overall spectrum of the emitted light. These non-linear effects typically include self-phase modulation, four-wave mixing, Raman scattering, and solitons dynamics, among others [14,17]. The propagation of these effects is performed by dispersing the pulsed laser beam through a non-linear material. These can range from bulk materials, such as simple water cells, silicas, or gases (as in the earliest attempts of SC generation), to more advanced crystal-structured optical fibres [17,18].
Early SC lasers using bulk materials for non-linear propagation were limited in the efficiency of coupling the input laser light with these propagating material (coupling efficiencies stood at below 10%) and thus they were able to only achieve ranges of about 200–300 nm, with pulses duration in the order of tens of nanoseconds. Further developments in SC generation efficiency and coupling using optic fibres lead to enlarging the range up to 1–2 μm and increasing the pulse duration to the range of picoseconds (for commercial, compact lasers), or even femtoseconds (for research-grade, benchtop lasers) [18]. A critical turning point for modern SC lasers to achieve such high performances was the invention and application of photonic crystal fibres (PCF), a particular type of microstructured optic fibre (Figure 1).
PCFs, invented in 1996 [19], are currently the most used non-linear propagating materials for SC generation [17,18]. As the name suggests, this class of optic fibres is based on recreating the optical properties of crystalline structures, in particular their photonic bandgap effect, by combining various high and low refractive indexes in their inner structure. Such features can be achieved in different manners, which typically define the given category of PCFs [20]: (1) hollow-core PCFs present a central hole in their structures, filled with air; (2) holey PCFs are composed of a matrix of several air-filled holes in their cross-sections that is interspersed with the core material; (3) solid-core PCFs involves a solid core surrounded by a matrix of air-filled holes.
PCFs offer high coupling efficiencies and low power losses, thus allowing ultra-broadband, high-brightness SC generation at cost-effective access [17,21]. SC lasers based on PCFs used in biomedical optics applications can typically generate broad spectrum spanning from 350 to 2400 nm, generally showing a central sharp peak which is a residual of the original shape of the pulsed beam. An example of this is shown in Figure 1. Most commercial SC lasers show performances similar to pulsed laser diodes, with integrated output power up to tens of watts and the capability of generating pulses in the order of picoseconds and repetition rates of tens of MHz [22].
Finally, we can report that in the recent year, more companies are able to provide ready-to-use SC solution. A search from the website https://www.rp-photonics.com (accessed date: 1 November 2021), displays at least 4 companies manufacturing SCL: NKT photonics, who recently acquired the former UK company Fianium (www.nktphotonics.com, accessed date: 1 November 2021), LEUKOS (www.leukos-systems.com, accessed date: 1 November 2021), FYLA (www.fyla.com, accessed date: 1 November 2021) and YSL photonics (www.yslphotonics.com, accessed date: 1 November 2021). Figure 1 displays pictures of different SCL offered by these companies.

3. The Use of Supercontinuum Lasers in Biomedical Diffuse Optics

The focus of the present review was to evaluate the use of SCL in biomedical diffuse optics, from the systems and methodological developments to their medical applications. Therefore, papers were identified using PubMed, Scopus, and Web of Science, searching for a combination of keywords including (supercontinuum laser|white laser) and (diffuse optics|NIRS). Moreover, a manual search from articles’ references was performed and the personal reference library of the authors of the current papers has also been searched. Papers were rejected if no SCL was used in the method section of the reported work, if the application was not medically related, if they were reviews or if the method was not purely optical (i.e., no photoacoustic papers). A total of 95 publications covering a variety of focus were identified. We have identified 4 main categories of publications: system development (31 studies), optical properties estimation (16 studies), methodology (43 studies) and application (5 studies). Table 1 provides a summary of the main characteristics of the SCL used in the studies reported in this review, Figure 2 shows pictures of the instrumentation typically used, and Figure 3 presents typical output in terms of datatypes of such systems. Moreover, a summary of the studies included in this review is presented in Table 2 and includes the system used, the year, the category of the study and the target of the study. Finally, Figure 4 summarizes the main topics and characteristic of the systems used in the reviewed publications and shows the number of publications over the last 30 years that report the use of SCL in biomedical diffuse optics.

3.1. Novel Instrument Developement

The first paper reporting the use of a supercontinuum laser in biomedical diffuse optics was reported in the early nineties [29,30,31] by a group based at the Lund Laser Centre (LLC, Lund, Sweden). The femtosecond white light supercontinuum was generated by a Terawatt laser running at 10 Hz. The spectral decomposition was performed on the detection side, by coupling an imaging spectrometer and a streak camera, enabling to acquire parallelly the DTOF at all wavelengths. To improve the signal-to-noise ratio, the measurement had to be repeated for several laser shots, but the overall acquisition time was of just a few minutes, which is acceptable when no dynamic contrast is targeted. Using such a system, the very first in vivo absorption and scattering spectra of the female breast were acquired from 650 to 850 nm. This system was then updated in the early 2000s with a supercontinuum generation system based on a crystal fibre, which allowed to increase the optical power of the SCL generation [32].
Parallel to these initial developments, a team based at Politecnico di Milano (POLIMI) produced a system able to characterise tissue between 610 and 1050 nm [33]. The system was however not based on a supercontinuum laser generation but on two tuneable lasers for the source, and the detection scheme was based on time-correlated single-photon counting (TCSPC), enabling light detection down to the single photon, which drastically improves the sensitivity of the system. This increased sensitivity is of great advantage to reduce the acquisition time and improve the light collection in the region where water absorption is high (900–1050). However, handling two lasers drastically increases the complexity and bulkiness of the system. Nevertheless, with that system, that team was able to record the optical properties of absorption and scattering of various tissues in-vivo [34,35,36,37]. The availability of commercial SCL sources gave rise to a new version of that system, where the 2 lasers were replaced by a SCL source coupled with a prism for the spectral selection [38]. This new source allowed to make the system more compact and transportable making it more suitable for the clinic. Moreover, the detector was also replaced (previously photomultiplier (PMT)) by a Single-Photon Avalanche Diode (SPAD) which enable to increase the spectral coverage, allowing a measurement between 600 and 1000 nm. This system could be used to characterise the optical and physiological properties of the tissue of the female breast. Another version of that system was developed, combining the SCL with a Ti-Sapphire laser to expand the spectral coverage to 1100 nm [39]. The SCL was used to cover the 600–900 nm and the Ti-Sapphire was used between 900 and 1100 nm. Indeed, even though the SCL could cover the entire considered wavelength band, the spectral resolution provided by the couple SCL/dispersion prism was too low to provide accurate results, as the source bandwidth increased with wavelength from 3 nm at 600 nm to 10 nm at 900 nm, and to approximately 15 nm at 1100 nm with that solution. This instrument has indeed been used to demonstrate that the broadening of the spectral bandwidth effect could impair the results, as the spectral region within the bandpass that exhibits the lowest absorption will dominate, leading to significant underestimations of the absorption and distortions in the spectral shape [40]. The spectral quality of the acquired data being extremely important, having the narrowest bandwidth is a great advantage. To do so, using a SCL is thus a true advantage as it provides a high power across the entire spectra. However, the method used to filter the white light is very important. As it has been seen above, several techniques can be used, either filtering on the detection side with a spectrometer, or on the source side with a dispersive prism or using an acousto-optic tuneable filter (AOTF), as we will see in systems described below. The POLIMI team has compared the wavelength selection from a SCL with an AOTF and a prism. They have shown that one had to be careful when using AOTF, as they can produce side lobes in the selected bandpass, which can drastically impair the retrieval of the optical properties [41].
Nevertheless, the extension of the spectral coverage offered by this newly developed system was of particular interest, as several components, like lipid and collagen, have a distinct signature above 1000 nm. The LLC group also reported a similar system, based on a commercial SCL coupled with 2 AOTF and a detection scheme based on TCSPC with a coverage between 650 and 1400 nm [42]. The POLIMI group has then extended this work by focusing on the spectral window between 900 and 1700 nm, by using an InGaAs/InP SPAD as a detector [43,44,45]. Finally, they have produced a system able to cover the entire bandwidth from 600 to 1350 nm using a single SCL coupled with a prism for the wavelength selection, and 2 TCSPC cards couples with 2 detectors covering different bandwidth [23]. They managed to fix the issue of the spectral broadening induced by the prism by refining the optical chain. Additionally, this system was mounted on a 19-inch medical cart, making it easily used in a clinical environment. This system can be seen in Figure 2 and in an example of data in Figure 3. Moreover, by adjusting the detector used, the same instrument can be used to cover an even larger bandwidth, between 500 to 1700 nm [46].
With all these developments, POLIMI has been a major contributor to multispectral TD-NIRS and were able to characterize various tissue over a large bandwidth. They were able to characterize bones [47,48,49], thyroid tissues [50], adipose tissues [51], collagen [46,52], and elastin [53]. This groundwork was crucial in order to investigate the tissue composition of breast for example, which can help for diagnosis of cancerous tissue [54,55]. Finally, we can report that the latest reported system has also been used to monitor the thermal effect of Radio Frequency (RF) on tissues [56,57].
The previously reported systems were only composed of one channel, which can be a drawback when an image is required, like for performing mammography. Therefore, parallel to these developments, another system aimed at optical mammography was developed, based on a supercontinuum light generated in PCF and a 32-channel parallel detection [58]. This allowed to acquired image of the breast between 606 and 885 nm, and to extract hemoglobin, tissue saturation, lipid and water content, and scattering parameters of the tissue.
All the systems reported above were designed to collect spectral information about tissue, with a first application in breast cancer detection. Therefore, the acquisition time was not limited, and functional contrast was not explored. However, the potential of monitoring real time change in optical properties over a large bandwidth is quite appealing for various application, like brain monitoring [59]. Therefore, other systems were developed in order to achieve that goal. A system was developed at POLIMI based on a supercontinuum source generated in a PCF by a self-locking mode-locked Ti:Sapphire oscillator and a detection scheme based on an imaging spectrometer coupled to a 16-channel multianode PMT connected to a TCSPC board [60]. The fact that all the 16 wavelengths were captured at once enabled a fast acquisition and permitted to track dynamic variations in absorption and scattering spectra following hemodynamic changes [61]. The limitation of the detections scheme constraint the bandwidth of this system between 520 and 850 nm.
A similar approach was taken by a group based in Poland [62,63]. The system was based on a commercial SCL for the source, coupled to a 16-channel multianode PMT connected to a TCSPC board for the detection part. This system was used to follow a contrast agent (Indocyanine green (ICG)), which allows to estimate the cerebral blood perfusion. The same system can be used without ICG to monitor the intrinsic optical contrast [64]. Similarly, the author and colleagues used a TD-NIRS system based on a SCL filtered at 808 nm and on a detection scheme based on a single channel TCSPC to follow ICG bolus [65]. The novelty of this study was the fact that the TD-NIRS was coupled with a diffuse correlation spectroscopy system (DCS) [66]. The coupling of these two techniques allowed to retrieve absolute cerebral blood flow measurement in patients in the intensive care unit, showing that these systems could be used in a particularly difficult clinical environment. Moreover, the possibility to also measure blood perfusion is of great interest and might help in the adoption of the technique in the clinic.
Another potential detection scheme in order to collect parallel data is to use an Intensified CCD (ICCD) camera. Selb and collaborators used a commercial SCL combined with a ICCD camera to perform functional brain imaging with 2 wavelengths at a frequency of 3 Hz [67]. This system was an upgrade of a previously developed instrument based on pulsed Ti:Sapphire laser [68]. This source was tuneable but the switching time between wavelength was too slow to enable dual-wavelength functional imaging. Therefore, the authors report that the use of SCL unlocked this possibility with their detection scheme.
Similarly, Lange and collaborators described a TD-NIRS system developed to measure the human brain tissue physiology [69,70]. This system was based on commercial SCL for the emission and on an ICCD camera coupled with an imaging spectrometer in order to perform the spectral decomposition. This system allowed to retrieve the typical hemodynamic response, and preliminary results on the detection of oxCCO were also performed. The ability to retrieve oxCCO is exiting for brain applications, as it has been shown that this contrast can be more specific than the oxygenation, and that it can be an early biomarker of neonatal encephalopathy for example [7]. This possibility has been explored by the UCL team by developing a 4-channel 16 wavelengths TD-NIRS system in order to monitor both the oxygenation and oxCCO signal [24,71]. This system was based on time multiplexing, with a TCSPC detection scheme, but the fast-switching capabilities of the SCL coupled with AOTF allowed the acquisition of all the 16 wavelengths in less than 2 s. This system has also been used to evaluate the long-term reproducibility of the evaluation of cerebral tissue saturation in healthy adults with good reproducibility [72]. This system can be seen in Figure 2, and in an example of data in Figure 3.
Another approach that has been used using SCL sources is the development of non-contact systems. This can present several advantages to avoid issues attributed to sensor-tissue contact like skin compression and to obtain a reduction in measurement preparation time. A team at the Physikalisch-Technische Bundesanstalt (PTB) has developed a non-contact TD-NIRS system based on a commercial SCL and fast-gated SPAD to explore the human brain physiology [73,74,75,76]. At the time, the authors reported that the use of SCL permitted a shorter pulse width (<100 ps), an achievable output power considerably larger than picosecond diode lasers, and the possibility to quickly switch between two different wavelengths (i.e., on the µs to ms time scale). Moreover, the use of the SPAD in that system enables to reject the burst of early photons, which can improve the spatial resolution of the system at very short detector distances [77]. This process had been previously demonstrated by the POLIMI team that had develop a single channel TD-NIRS system based on SCL and a fast gated SPAD detector [78]. The POLIMI team has also worked on the modelling of the parameters of such detection schemes, in order to optimise their design [79].
The non-contact scheme is also very useful for pre-clinical applications. Indeed, several systems, based on SCL for the source, aiming at collecting physiological information of small animals, have been proposed. In the early 2000s, a system based on a home-made SC generation associated with an imaging spectrometer coupled with a streak camera was developed and used to monitor the cerebral hemodynamic of songbirds [80,81]. More recently, the UCL team have developed a system aiming at investigating both the cerebral hemodynamic and metabolic response (via oxCCO) of the exposed cortex of rodents [25]. In this work, a commercial SCL was used in combination with a dispersive prism, for the emission, and a sCMOS camera, as a detector, to record 11 images (11 wavelengths) of the exposed cortex in less than 0.3 s/wavelength. Here the recording was not time resolved, but the power of the laser and the ability to quickly switch between the wavelength had dictated the choice of this source. This system can be seen in Figure 2 and an example of data in Figure 3. Another approach can be to use spatial frequency domain imaging (SFDI), a wide-field, noncontact, model-based technique that can quantitatively assess absorption and scattering, on a pixel-by-pixel basis by calculating the modulation transfer function of structured light projected onto the tissue. This technique can be hyperspectral if a proper source is used. However, the system developed thus far was based on conventional broadband sources coupled with tunable filters, which limited their acquisition speed and spectral resolution. Torabzadeh and colleagues [82] applied this technique using the principles of the single-pixel compressive sensing. They used a commercial SCL coupled with a dispersive prism to achieve a final bandwidth 1000 spectral bands between 580 and 950 nm, which was larger than the previously developed systems. With this system, the authors were able to characterize the optical properties of a beef sample ex-vivo in 2D, and extract physiological information (i.e., hemoglobin content and fat and water content). The authors noted that the wavelength selection technique used was still slow (i.e., 150 s of acquisition time for a 4 cm × 6 cm), but that switching to an AOTF solution could greatly improve this speed. Using the same principle of the single-pixel compressive sensing, Farina and colleagues reported a TD-DOT instrument based on time-resolved single pixel camera [83]. Their approach was based on a SCL source filtered at 620 nm coupled with a digital micromirror device (DMD) to generate the spatial pattern. The time domain acquisition was based on TCSPC. Pian and colleagues used a similar approach but expended it to a broadband illumination and detection [84]. The authors reported that their system, aimed at fluorescence molecular tomography, benefitted for the commercial SCL source to perform hyperspectral excitation of the sample which offers a spectral information boost within the same data acquisition time compared to single wavelength excitation situations.
Another work aiming at improving the spatial resolution of TD-DOT was to investigate the effect of only considering early photons, which undergo fewer scattering events [85]. In this work, this approach has been tested using a commercial SCL together with a bandpass filter at 670 nm. It worth noting that the commercial version of this laser was modified by the manufacturer so that it had a shortened fibre compared to standard models, which reduced the pulse width to approximately 300 ps. This is important as the benefit of the gating of the early photons is highly dependent on the instrument response function (IRF) of the system. The authors reported that with this technique, and this instrument, a 64 to 84% image resolution could be achieved, which could be a great benefit for preclinical imaging with diffuse optics.
Novel algorithms can also be used in combination to TD-DOT instruments to recovers 3D images of the optical properties and physiological parameters. A team in Grenoble has explored the use of the Mellin-Laplace transform to reconstruct 3D images of the chromophores inside phantoms and for the non-invasive assessment of flap viability in rats’ models. There instrumentation was based on commercial SCL, coupled with either an AOTF or an interferential filter, for the source and a TCPSC scheme for the detection [86,87].
TD-DOT has also been explored in humans, in order to reconstruct 2D or 3D maps of the optical properties and hemodynamic parameters. To this end, the UCL team had design MONSTIR II to perform tomographic brain imaging [28]. This system is based on a commercial SCL with an AOTF filtering and has 32 channels based on TCSPC. It was successfully used to scan healthy and ill infants in a clinical environment [88,89], an example of data produced by this system is provided in Figure 3. More recently, a new TD-DOT system based on a commercial SCL filtered by an AOTF, and a 32 × 32-pixel SPAD camera, has been developed by a team at the Biomedical Optics Research Laboratory at the University of Zurich [26,27,90]. This system operates at 2 wavelengths and at 11 sources points, which can be scanned in about 3 s. It is a big step forward compared to previously developed system, especially considering its size, as it is handheld. A picture of this instrument is reported in Figure 2.
Finally, all the instrument mentioned above were mostly based on the TD technique, or traditional white light illumination. We can note that we found one frequency domain (FD) system, that is based a commercial SCL to collect our system collects FD data at 170 wavelengths from 680 to 850 nm and allows to accurately measure absorption and scattering information about the tissue [91].

3.2. Methodology

We have seen that the SCL has been used to develop several systems, mainly aiming at characterising the optical properties of tissues over a large bandwidth or to non-invasively follow the oxygenation and/or metabolic changes both in humans and in preclinical models. Both of these capacities are extremely appealing for clinical applications. However, the transition to the clinical application of system based on SCL and more generally TD-NIRS/DOT is still happening at a slow pace. One of the reasons of the slow adoption of optical techniques is the lack of standards in the community. Indeed, standardisation is key in order to test all the systems on the same ground, and reenforce the trustworthiness of the technique. In the recent years, standardisation have been a main goal of the NIRS community, and several protocols have been developed in order to compare and test the capabilities of NIRS systems. The main produced protocols were the MEDPHOT [92], BIP [93], and nEUROpt [94]. It is out of the scope of this paper to describe these protocols in detail, but the cornerstones of all of these are to produce standard phantoms, well characterized spectrally in terms of absorption and scattering coefficients, in order to be able to quantitatively compare different systems or methods. In this matter, SCL based systems were extensively used to characterize the material used to produce these phantoms and the actual final phantoms used as standard. These optical phantoms can either be solid (easier to store and stable over a long period of time) or liquid (easier to precisely tune the optical properties of the solution and to adjust it dynamically). For the liquid phantoms, the core principle is to use a fatty solution to control the scattering coefficient of the solution, and ink to control their absorption properties. Several recipes have been tested, like using Agar and Triton [95], but the most common recipe to date is to use a combination of water, intralipid and Indian ink. This choice has arisen after a careful characterization of both intralipid and Indian ink, principally with instrument based on SCL [96,97,98,99,100]. We can note that for the precise optical characterisation of liquid, techniques based on integrating sphere are often used. However, this methodology cannot be applied for large solid phantoms. In this matter, TD-NIRS has proved a good methodology to accurately estimates the optical properties. For example, in 2010, Bouchard and colleagues used a TD-NIRS setup based on a SCL source filtered at 610 nm in order to estimate the optical properties of phantom with an absolute error estimates of 0.01 cm−1 (11.3%) and 0.67 cm−1 (6.8%) for the absorption coefficient and reduced scattering coefficient, respectively [101].
On top of designing phantom able to mimic the background optical properties of tissues, some methods have emerged to mimic a perturbation of the absorption coefficient within a homogeneous medium. To that end, one can either use a liquid phantom for the homogeneous medium, and a small black PVC target for the perturbation. A protocol has been designed and it is possible to precisely set the absorption perturbation value based on the size of the black PVC inclusion [94]. This protocol has been validated with Monte Carlo simulations and in-situ using a system based on SCL [102]. A similar phantom based on a completely solid phantom, with a movable solid rod, has also been designed and characterised [103]. Lastly, some phantoms have also been characterized in order to be a tool to evaluate quantitatively the responsivity of NIRS systems [104].
Finally, we can also mention new approaches based on digital phantoms. Basically, rather than physically simulating the process of light propagation in tissue, the digital phantom provides the detector with light signals mimicking the signals obtained in-vivo. Such systems based on an integrating sphere and several light sources, including a commercial SCL, have been proposed to spectrally characterise and calibration optical systems [105]. More recently, a digital phantom has been developed to mimic TD-NIRS signal acquisition, and the authors used various sources and detectors in order to test this approach, including a SCL [106]. One other important component to optically characterize is the optode holders used to secure the optical fibres to the patient. In a recent study, Amendola and colleagues used a system previously developed to evaluate the optical properties of fruit [107] (based on a commercial SCL and a filter wheel to select 14 spectral band in the range 550 to 940 nm) to spectrally characterised poly lactic acid (PLA), a material used for 3D printing [108]. As 3D printing is more and more common in order to produce probe holders, it is important to optically characterise the material used in order to avoid any unwanted effects on the measurement provoked by light interaction with the material. In that study, the authors concluded that the different material tested were not optically the same, and that characterisation was important before their use.
The development of all these phantoms has provided a good framework for the developers to compare their systems and methods in order to provide robust clinical information. Regarding the instrumentation, a lot of effort are currently being made in order to improve the detection scheme of TD-NIRS. To that end, SCL have been a source of choice to test new detectors like the promising fast silicon photomultiplier (SiPM) which drastically improve light harvesting compared to the old PMTs [109,110,111]. SCL based systems have also been used to characterise novel bio-compatible fibres that can be implanted in the patient with potential use in the monitoring of the evolution of the physiology after a surgical intervention or in photodynamic therapy (PDT) [112,113].
Regarding the methodology, several new data processing techniques have been explored and tested in conjunction with SCL. Indeed, system based on SCL can collect multidimensional data that can be coupled together in order to improve the accuracy of the data, reduce the noise or explore new contrast. One of the first work regarding this matter dates from 2006, where D’Andrea and colleagues, showed that one could increase the robustness of the determination of the chromophore’s concentrations in tissues (hemoglobin, water, lipid) by using a spectral constraint during the fitting procedure [114]. This showed the strength of using the spectral dimension coupled with the DTOF, which is accessible when using SCL. This method has also been used to monitor absorption changes in a layered diffusive medium [115], which is relevant in the monitoring of brain activation for example. Similar approaches were explored in the same group in order to accurately monitor the spectral changes in absorption coefficients of turbid media [116].
One can also use the spatial dimension in order to enhance the amount of information recorded, by recording data from multiple source-detector distances. This approach has been recently tested and it has been shown that coupling the spatial dimension with the DTOF could increase the accuracy of the estimation of the optical properties [117,118,119]. Another interesting use of the spatial dimension is the self-calibrating method for TD-NIRS, which uses the spatial dimension in order to avoid the measurement of the IRF, that is needed to remove the effect of the instrument from the response of the tissue [120]. The characterisation of the IRF necessitates an extra step that can be a burden in a clinical environment. Therefore, this method could facilitate the use of TD-NIRS in the clinic. Finally, we can also report that an approach combining the spatial and spectral dimension proved to also improve the accuracy of the results of CW-NIRS. This has been tested recently, especially with blood phantoms and compared to the results of a TD-NIRS instrument based on a SCL source [121].
The large amount of data collected by TD-NIRS also enables different data processing scheme to extract the relevant information. Therefore, there is a need to properly compare all the available methods in order to evaluate their strength and weaknesses. This work has recently been undertaken in order to compare the moment and time windowing techniques [122,123]. Having a deep understanding of all the data processing method is also crucial so the proper technique can be used depending on the application.
Several multi-centre initiatives have recently promoted this standardisation effort to compare instruments and algorithms. To that end, partners from the BitMap network (http://www.bitmap-itn.eu/, accessed date: 1 February 2021) have started comparing instruments from different institutions all over Europe, by using the same sets of solid phantoms from the 3 main protocols [124,125]. Moreover, several centres in Europe have used their instruments (several based on SCL) on 9 subjects, to study the inter-subject variability of the optical properties of the human head measured by several instruments [126]. They could show that the inter-subject variability was significant, with a big effect of the technique used (either CW or TD) because of their different depth sensitivity, which in turn affects the averaged optical properties retrieved. Therefore, it is evident that all the efforts to precisely characterized instruments and methods will have a huge impact on the precision of the optical measurement of tissue parameters.
Finally, the use of SCL can help to explore new avenues in diffuse optics. The main one is to go above the “therapeutical window” in the second (1100 nm to 1350 nm) and the third (1600 to 1870 nm) near-infrared optical window. On top of the previously reported works that expand the bandwidth of TD-NIRS systems up to 1700 nm, Sordillo and colleagues reported the use of a commercial SCL source, filtered with a bandpass and long pass filters to target the second and third NIR windows, coupled with an IR-CCD camera, to acquire transmission images of 3 targets embedded in chicken tissue of various thickness [127]. They were able to distinguish the targets up to a depth of 10 mm, which was not possible when using the same setup but with a traditional lamp source. These results are promising in order to implement transmission measurement, especially for preclinical imaging.
As we have seen previously, the use of these long NIR wavelength is useful to resolve more chromophores like collagen. It is also promising to resolve glucose, which could be a true benefit for several clinical applications like monitoring diabetes [128], or monitoring brain injuries [129] or neonatal encephalopathy [130]. Once again, in this wavelength range, the power delivered by SCL has proved efficient to monitor the glucose contrast in the NIR [131,132].
Finally, a tentative to non-invasively probe the lungs with a TD-NIRS system based on SCL source was reported [133]. The measurements were performed on three subjects, and spectra from 600 to 1100 nm could be acquired. Here, the depth sensitivity of a TD-NIRS is essential and this application shows that instruments based on that technology could be used to non-invasively probes tissues previously thought out of reach.
Table 2. Table of the literature review of the use of SC lasers in biomedical diffuse optics. The system used in each study are described more in detail in Table 1 and referred to as “system ID”. µa: absorption coefficient, µ’s: reduced scattering coefficient.
Table 2. Table of the literature review of the use of SC lasers in biomedical diffuse optics. The system used in each study are described more in detail in Table 1 and referred to as “system ID”. µa: absorption coefficient, µ’s: reduced scattering coefficient.
PublicationYearSystem IDCategory of StudyTarget of StudyReported Quantities
S. Andersson-Engels et al. [29]19931System developmentMultispectral tissue characterisationµa
af Klinteberg et al. [30]19951Optical properties estimationBreast tissue examinationµa, µ’s
Bassi et al. [60]20042System developmentPhantoms and in vivo validation measurementsµa, µ´s
Abrahamsson et al. [32]20043System developmentSample characterisationµa, µ´s
Swartling et al. [61]20053System developmentPhantoms and in vivo validation measurementsµa, µ´s, [HbT], StO2
Ramstein et al. [81]20054System developmentOptical properties in vivo monitoring of songbird brainµa, µ´s
D’Andrea et al. [114]20062MethodologyValidation of spectral fitting analysisµa, µ´s, H2O, Water Content Lipid Content, [HbO2], [HHb], [HbT], StO2
Bassi et al. [58]20062System developmentOptical mammographyµa, µ´s, H2O, Lipid Content, [HbT], StO2
Bassi et al. [38]20076System developmentPhantoms and in vivo breast validationµa, µ´s, H2O, Lipid Content, [HbO2], [HHb], [HbT], StO2
Vignal et al. [134]20084ApplicationIn vivo measurement of brain hemodynamic changes in songbird[HbO2], [HHb], [HbT]
Farina et al. [40]20096MethodologyStudy and correction of bandpass effects in TD spectroscopyµa
Taroni et al. [39]20096System developmentIn vivo measurements of breast tissueµa, µ´s, Water Content, Lipid Content, Collagen Content, [HbT], StO2
Svensson et al. [42]20097System developmentValidation on phantomsµa, µ´s
Valim et al. [85]20108MethodologyStudy of PDSF in phantomsPhoton density sensitivity function (PDSF)
Bouchard et al. [101]20109MethodologyCharacterisation of tissue-mimicking phantomsn, g, µa, µ´s
Pifferi et al. [116]201015MethodologyData processingΔµa
Taroni et al. [52]201015Optical properties estimationCharacterisation of collagen optical propertiesµa, µ´s
Giusto et al. [115]20102MethodologyData Processingµa, µ´s, [HbO2], [HHb]
Dalla Mora et al. [78]20106MethodologyElectronic developmentµa, µ´s
Mottin et al. [80]20114ApplicationIn vivo study of brain oxygen uncoupling/recoupling in songbirds[HbO2], [HHb]
Mazurenka et al. [73]201110System developmentNon-contact TD-NIRSContrast
Arnesano et al. [91]201211System developmentFD spectroscopy for tissue imagingµa, µ´s
Spinelli et al. [97]2012Multiple systemsOptical properties estimationCharacterisation of liquid phantomsIntrinsic absorption coefficient, Intrinsic reduced scattering coefficient
Bargigia et al. [44]201213System developmentCharacterisation on lipidsµa, µ´s
Gerega et al. [63]201212MethodologyTD-NIRS with ICGFluorescence of ICG
Bargigia et al. [43]201215System developmentBandwidth up to 1700 nmµa, µ´s
Xu et al. [105]2012Laser not specifiedMethodologyDigital phantomStO2
Wang et al. [96]201216System developmentLiquid optical phantom characterisation in the second and third optical windowg, µa, µ´s
Selb et al. [67]201314System developmentFunctional brain imagingΔ[HbO2], Δ[HHb]
Bargigia et al. [45]201313System developmentIn vivo measurements of forearm and breastWater Content, Lipid Content, Collagen Content
Farina et al. [41]20136MethodologyComparison of approaches for spectral selectionµa, µ´s
Mazurenka et al. [75]201326System developmentNon-contact systemΔ[HbO2], Δ[HHb]
Wabnitz et al. [135]2013Multiple instrumentsMethodologymulti-laboratory study to assess the performance of time-domain optical brain imagersµa, µ´s
Quarto et al. [95]201315MethodologyPhantom characterisationµa, µ´s
Aernouts et al. [98]201316Optical properties estimationPhantom characterisationg, µa, µ´s
Quarto et al. [133]201315MethodologyIn vivo optical diagnostics of lung conditions and diseasesµa, µ´s
Aernouts et al. [99]201416Optical properties estimationPhantom characterisationg, µa, µ´s
Spinelli et al. [100]20146Optical properties estimationPhantom characterisationµa, µ´s
Cooper et al. [28]201417System developmentBrain TD-DOTµa, µ´s
Farina et al. [126]2014Multiple instrumentsOptical properties estimationBrain Tissue characterisationµa, µ´s
Wabnitz et al. [93]201417MethodologyInstrumental performance protocolInstrument parameters
Wabnitz et al. [94]201417MethodologyInstrumental performance protocolInstrument parameters, µa
Martelli et al. [102]201418MethodologyPhantom characterizationµa, µ´s
Taroni et al. [55]20156Optical properties estimationIn vivo quantification of collagen in breast tissueWater Content, Lipid Content, Collagen Content, Collagen index, [HbT], StO2
Lange et al. [70]201519System developmentFunctional brain monitoringΔ[HbO2], Δ[HHb], Δ[oxCCO]
Farina et al. [83]201520System developmentTD-DOT/FMTFluorescence
Contini et al. [79]201521MethodologyTime-gated measurements on tissue phantomsContrast
Della Mora et al. [109]201522MethodologyTesting of a SiPMµa, µ´s
Pifferi et al. [103]201515MethodologyPhantom characterisationµa, µ´s
Martinenghi et al. [110]201515MethodologyCharacterisation of SiPM detectorsInstrument parameters
Dempsey et al. [88]201517MethodologyWhole-head TD-DOT in neonatesµa, µ´s
Sordillo et al. [127]201523MethodologyDeep-tissue, optical properties monitoringµt
Konugolu Venkata Sekar et al. [47]201524Optical properties estimationIn vivo measurement of optical properties of boneµa, µ´s, Water Content, Lipid Content, Collagen Content, [HbO2], [HHb], [HbT], StO2
Konugolu Venkata Sekar et al. [23]201624Optical properties estimationIn vivo human tissues measurementsµa, µ´s, H2O, Lipid, Collagen, [HbO2], [HHb], [HbT], StO2
Di Sieno et al. [76]201625System developmentValidation on phantomsInstrument parameters
Wabnitz et al. [104]201626Optical properties estimationPhantoms characterisationµa, µ´s
Konugolu Venkata Sekar et al. [49]201624Optical properties estimationIn vivo measurement of optical properties of boneµa, µ´s, Water Content, Lipid Content, Collagen Content, [HbO2], [HHb], StO2
Konugolu Venkata Sekar et al. [48]201624Optical properties estimationIn vivo measurement of optical properties of boneµa, µ´s, Water Content, Lipid Content, Collagen Content, [HbO2], [HHb], [HbT], StO2
Martinenghi et al. [111]201613MethodologyCharacterisation of SiPM detectorsInstrument parameters
Di Sieno et al. [87]201627System developmentIn vivo assessment of flap viability[HbO2], [HHb]
Zouaoui et al. [86]201722MethodologyValidation of chromophore decomposition algorithmDyes content
Konugolu Venkata Sekar et al. [46]201724Optical properties estimationIn vivo measurement of optical properties of collagenµa Collagen
Wabnitz et al.[74]201727MethodologyNon-contact TD brain imagingIntensity Contrast
Di Sieno et al. [112]201724MethodologyCharacterisation of bioresorbable fibresµa, µ´s
Konugolu Venkata Sekar et al. [53]201724Optical properties estimationIn vivo measurement of optical properties of elastinµa and µ´s Elastin
Jiang et al. [90]201739MethodologyData Processing—GPUµa, µ´s
Konugolu Venkata Sekar et al. [50]201824Optical properties estimationIn vivo chromophore characterisation of thyroidµa, µ´s
Lange et al. [69]201819System developmentIn vivo monitoring of brain physiological changes in humansΔ[HbO2], Δ[HHb]
Pian et al. [84]201829System developmentValidation on phantomsµa
Gerega et al. [62]201830Applicationassessment of intracerebral and extracerebral absorption changesΔ[ICG]
He et al. [65]201831Applicationcontinuous monitoring of absolute cerebral blood flow (CBF) in adult human patientsCBF
Della Mora et al. [113]201815Methodologybioresorbable optical fibersContrast, µa, µ´s
Laura Dempsey [89]201817System developmentBrain TD-DOT[HbO2], [HHb]
Liu et al. [131]201828MethodologyGlucose monitoring[Glucose]
Wabnitz et al. [106]201932MethodologyValidation of digital phantomPhantom parameters and basic responses of the instrument to it
Fuglerud et al. [132]201933MethodologyGlucose sensing in blood[Glucose]
Torabzadeh et al. [82]201934System developmentSpatial FD HIS on ex vivo beef sampleµa, µ´s, Water Content, lipid Content, [HbO2], [HHb], [MHb]
Lange et al. [24]201935System developmentValidation on phantomsµa, µ´s, [HbO2], [HHb], [oxCCO]
Ferocino et al. [54]201924MethodologyValidation on phantoms of TD-NIRs reconstruction algorithmµa, µ´s
Lange et al. [72]201935ApplicationReproducibility analysis of cerebral oxygenation measured with TD-NIRS[HbO2], [HHb], StO2
Yang et al. [118]201936MethodologyValidation on phantoms of a multivariate TD-SD algorithmµa, µ´s
Yang et al. [117]201936MethodologyValidation on phantoms of a multivariate TD-SD algorithmµa, µ´s
Ferocino et al. [136]201924MethodologyValidation on meat samples of TD-NIRs reconstruction algorithm[HbO2]
Wojtkiewicz et al. [120]201930MethodologyValidation on phantoms and in vivo of a self-calibrating TD algorithmµa, µ´s, [HbO2], [HHb]
Sudakou et al. [64]201930System developmentMulti wavelength TD-NIRS system for brain monitoring[HbO2], [HHb], [HbT]
Wabnitz et al. [122]202017MethodologyDepth-selective analysis in TD optical brain imagingµa, µ´s
Lanka et al. [51]202024Optical properties estimationIn vivo measurement of optical properties of adipose tissueµa, µ´s, Water Content, Lipid Content, Collagen Content, [HbO2], [HHb], [HbT], StO2
Sudakou et al. [123]202037MethodologyPerformance of measurands in TD optical brain imagingµa, µ´s
Yang et al. [119]202032MethodologySpace-enhanced TD diffuse optics in two-layered structuresµa, µ´s
Lanka et al. [56]202024MethodologyMonitoring of thermal treatment in biological tissuesµa, µ´s
Amendola et al. [108]202024MethodologyEffect of the 3d printed material on raw dataPhoton Counts
Jiang et al. [26]202039System developmentNovel TD-NIRS system based on SPAD cameraµa, µ´s
Jiang et al. [27]202039MethodologyImage reconstruction for TD-NIRS system based on SPAD systemµa
Kovacsova et al. [121]202135MethodologyValidation of broadband oximetry algorithmµa, µ´s, StO2
Giannoni et al. [25]202138System developmentHSI of in vivo hemoglobin and CCO in the exposed cortex of miceΔ[HbO2], Δ[HHb], Δ[oxCCO]

4. Discussion

We have summarized the use of SCL sources in biomedical diffuse optics. Historically, these sources were allowed to develop TD-NIRS system as they could deliver a pulsed coherent light at several MHz with narrow pulse width in order to reduce the IRF [137], and high power, of several mW/Wavelength, which were characteristically difficult to achieve in the early 1990s. Moreover, the development of commercially available compact SCL in the 2000s also helped the development of the use of these sources, and we have reported that more companies (see Figure 1) are now able to provide such sources.
The main advantage of these sources is that they make it easier to explore more wavelength, enabling a multispectral or even broadband or hyperspectral measurement. Several measurement strategies can be implemented in order to do so. The first one is to use the white light as a source and achieve the spectral unmixing at the detection stage. The first available option to do so was to use an imaging spectrometer coupled with a streak camera, which achieve very high temporal resolution (in terms of arrival time of photons). However, this solution was bulky, and the sensitivity of these camera was low compared to the newest detection technologies. The same method can be used by either coupling the spectrometer with an ICCD camera, or a 16-channel TCSPC. This approach of performing the spectral unmixing on the detection can achieve faster measurement since no spectral switching mechanism is employed. However, the resolution of these approaches, either spectrally or temporally, can be limited. For example, typical ICCD cameras have a gate width of 200 ps, which limits the temporal resolution of the DTOF. On the other end, the 16-channel TCSPC system used by Sudakou and colleagues [64] allows spectral bands of 12.5 nm, limiting the spectral resolution. The other option to perform the spectral unmixing is to do a sequential acquisition, by filtering the white light on the source side. To do so, two main approach are used, (1) using an AOTF, with the ability of fast switching (few microseconds) but with a potential degradation of the spectral response (i.e., side lobs), and (2), using a dispersing prism which can be mounted on a rotating axis, which can achieve better spectral performances but at the cost of a slower switching time. Therefore, the choice of the instrumentation depends on the end application, and trade-off needs to be made between spectral quality and acquisition speed. Whatever the requirement, we have seen that SCL can be used and coupled with various type of instrumentation in order to accommodate to the requirements of the applications.
From an application point of view, the ability to monitor a large number of several wavelengths has several advantages. First of all, it enables to provides a true spectroscopic analysis, both in terms of absorption and scattering. This is extremely appealing for cancer application, where this type of “optical biopsy”, enables to distinguish between healthy and cancerous tissue. This type of application has been extensively used in breast cancer [138,139], and more recently, thyroid applications [50,140]. In this type of application, where the spectral quality is the most important criteria, the second instrumental approach, based on spectral unmixing with a dispersing prism, was used. Secondly, the first instrumental approach, based on spectral unmixing at the detection stage, or the second instrumental approach using ATOF, give the possibility to acquire several wavelengths, either parallelly, or with a fast switching, in a time scale compatible with functional imaging. With these approaches, several authors have explicitly noted that the adoption of SCL enabled them to design or update existing systems, in order to retrieve dynamic physiological information [67,75].
The large number of wavelengths enabled by SCL also give the possibility to go beyond the traditional contrasts, like the hemodynamic one, in order to get information about metabolism for example, by extracting information about more chromophores, like oxCCO [7,24]. We can also note that beyond the intrinsic optical signal, a contrast agent, like ICG, can also be used in order to get information on tissue perfusion. This has notably been extensively used to monitor cerebral perfusion, and the SCL has also proved to be a suitable light source for these applications, even used in extremely complicated clinical settings like ICUs [65].
From a historical point of view, we can note that the development of commercial fibre also largely helped with the adoption of that technology. Indeed, out of the 35 different systems used in the reviewed paper, only 4 (less than 12%) are home-made. They correspond to the system developed before 2010. Past that date, all systems reviewed used a commercial source. We can also see the increase number of papers since the adoption of the commercial sources, with a number of papers published per year roughly doubling from that date. One of the main advantages reported by the authors when they adopted commercial SCL was their relative compactness, compared to the initial development of SC generation. This can be seen in Figure 2, with systems able to be mounted on trolley compatible with the clinical environment. However, the use of these lasers in actual clinical studies remains marginal. In our recent review [141], which has explored the use of TD-NIRS for clinical applications, less than 4% of the 52 papers reviewed were using a system based on a SCL source. Indeed, despite their advantages, a few drawbacks still hindered the adoption of SCL as a standard source. The first one is the cost which, despite a significant decrease in the recent years, remains high (i.e., more than 20,000 pounds for a SCL with several watts of power, enabling a final power of at least 1 mW per wavelength). The second main disadvantage is the warmup time, and the overall stability of these sources. Indeed, the stability of the output can be a concern, but the use of a reference arm can solve this issue [23]. This of course comes at the cost of a more complicated and bulkier instrument. Regarding the warmup time, typically, 1 h is required so the instrument reach thermal equilibrium, which can be too long for certain clinical applications. Some recent TD-NIRS systems based on pulsed laser diode have reported a warmup time of only 15 min, which is way more advantageous in a clinical setting [142]. Finally, even though the novel commercial SCL are more compact than in the early days, their size is still significantly bigger than traditional pulsed laser diode, which limit the reduction of the overall instrument size. Therefore, this can also slow down the adoption of these systems, as space are often very limited in busy clinical environment like ICUs.
These drawbacks are less of an issue for the preclinical settings as the environment is way less constrained. Therefore, the great potential of applications based on SCL outweighs its current main drawbacks, and several instruments based on SCL have been developed for preclinical imaging. The interested reader can refer to the recent review by Bérubé-Lauzière and colleagues, which explore more particularly the use of DOT in preclinical settings [143]. It worth noting that these systems are non-contact, and that this particular detection approach has also been explored for human imaging. Here, having an extremely narrow pulse, like the one provided by SCL, is of great interest as one can take advantage of the null-source approach [77]. Even if these approaches are less common, they can have potential great applications for perioperative measurements, or when the patient is difficult to access, and represent a promising avenue.
The above-mentioned drawbacks are also less problematic in a typical lab setting. SCL are a great tool for teams aiming at instrumental and methodological developments. In the reviewed papers, 33% were aimed at the initial system developments, and 18% and 44% were focused on optical tissue estimations or methodological developments, respectively. Only 5% of the reported studies were focused on pure applications, like tissue oxygenation estimation or cancer monitoring. Indeed, the availability of these SCL sources in teams focused on methodological and instrumental developments have pushed their used to test new possibilities and enhance the reliability of the system by promoting standardization. It is worth mentioning that having a SCL is not mandatory in a lot of settings, and that these sources were used as they were available in most of the groups working on instrumental and methodological developments. Indeed, in the case of a CW system, we have seen that SCL systems were used to provide a high number of wavelengths. In that case, other broadband sources are available, like Tungsten Halogen Light Sources (THLS) or like the Laser-Driven Light Sources (LDLS). The main disadvantage of the THLS is the low power density/wavelength, compared to coherent sources, which can limit their possibilities in biomedical diffuse optics. However, we can still report that THLS were successfully used to developed simple CW instruments [144,145]. On the contrary, the novel LDLS are of interest as they can produce a stable, high power broadband light. This technology is still relatively new and, to our knowledge, has not been used in biomedical diffuse optics yet. However, its attractive characteristics could be of interest for developing CW biomedical diffuse optics systems. The interested reader can refer to review [146] for more details. Finally, we can note that designing a multi-wavelengths CW system by adding up several single wavelengths light sources, like LEDs, is possible, but it is limited to a low number of wavelengths as the bulkiness and cost of the systems would rapidly increase with the number of wavelengths. In the case of TD systems, the choice of the light source is also largely dictated by the number of wavelengths needed, as can be seen in Table 3, that summarizes the relative advantages and disadvantages of single wavelength laser diode sources versus SCL in two cases: (1) low and (2) high number of wavelengths for TD systems. The use of single wavelength laser diode sources is indeed well adapted when a low number of wavelengths is required, as these sources are compact and relatively not expensive regarding the SCL ones. However, if one wants to design a truly spectroscopic system, with a high number of wavelengths required, these advantages became less evident as, similarly to the CW case, the bulkiness and cost of the system increase as function as the number of wavelengths. This is even more pronounced for the TD cases as the electronics are more complex. Finally, we can mention that if one wants to focus on methodological development, when specific wavelengths are still not defined, the flexibility of SCL is then evident as this choice does not have to be made from the start of the designing process, and not all the wavelengths are available commercially with laser diodes. Therefore, even though other light sources with characteristics compatible with diffuse optics are available, the high power, broadband spectrum, and fast switching capabilities of SCL makes them a source of choice, sufficiently versatile to be able to explore various avenues without source limitations.
This spectral flexibility offered by SCL has indeed been a true advantage regarding the standardization efforts, giving the ability to optically characterize over a large bandwidth material and tissue. Phantom materials, like Intralipid or Indian ink, have been able to be precisely characterized using SCL, enabling to produce reproducible recipes for liquid phantoms, or to produces reliable solid phantoms. This now gives the possibility to accurately and quantitatively test and compare newly developed instruments and algorithms. We can note the effort undertaken by the BitMap project that have started comparing instruments, from different institutions all over Europe, using the same sets of solid phantoms [124,125].
SCL can also be used as companions to develop new electronical components. Indeed, one of the big instrumental development in the recent year have been the refinement of the detection scheme. Especially, a lot of improvement have been seen in novel solid state detectors [147] which enable to boost light harvesting detectors electronics. The development of gated detector [148] also enables techniques like null source distance to be practically implemented, which can improve the resolution and sensitivity of the technique [77]. In these works, the availability of SCL sources enabled to perform a precise characterisation of these detectors in situ, both in terms of temporal resolution, thanks to the short pulse width of the SCL, and in terms of spectral responsivity over a large bandwidth. The combination of the SCL capabilities with these new detectors can notably boost the capabilities of TD-NIR instrument, overcoming some current limitations like, for example, the penetration depth [22,149].
The inherent ability of SCL sources to construct systems gathering multidimensional information, i.e., spatial/spectral/temporal, also makes them a perfect tool to test novel algorithms able to exploit this large amount of data. The recent works by Yang and colleagues [117,118,119] show the advantages of using multi-dimensional datasets, by enhancing the accuracy of the optical properties estimation of tissues by using simultaneously the temporal and spatial dimensions. Kovacsova and colleagues [121] also showed this utility, here combining the spectral and the spatial dimensions. It worth noting that this work is based on broadband CW-NIRS, but that the results of this algorithm has been tested against a TD-NIRS instrument based on SCL as a source. Combining all the dimensions available is certainly a really exiting avenue for the field of diffuse optics, to make the system even more robust and precise.
Finally, we can mention that in the recent years, progress has also been made in individual pulsed laser source, which now achieve several 10th of mW and a pulse width shorter than 100 ps. A good example of the uses of these sources can be found in [150], where an 8 wavelengths TD-NIRS system, compatible for breast imaging for example, has been developed. This paves the way on the use of these sources to enable miniaturisation of TD-NIRS instruments, making them truly compatible with the clinic. These sources could then replace the SCL in clinical settings. However, depending on the application and the chromophores targeted, the wavelength selection is a process that requires some investigations in order to balance the system complexity, and the possible crosstalk effects between the chosen wavelengths. Therefore, systems based on SCL able to easily explore various combinations of wavelengths are once more extremely useful. Lastly, the flexibility of these systems also allows to explore new horizons and applications area, like we have seen with exploring contrast above the classical therapeutic windows (i.e., above 1000 nm) [46], or to explore new organs, like lungs [133]. Therefore, even though the main drawbacks of SCL in terms of absolute cost, size, and stability might prevent their use on a wide scale in the clinical world, their main advantages in terms of power, wavelength choice, and relatively low cost for high number of wavelengths is likely to still push their use in the labs to drive forward the clinical innovation in biomedical diffuse optics.

5. Conclusions

We have seen that the use of SCL drove the developments in biomedical diffuse optics, and particularly in TD-NIRS. First of all, it is a tool of choice to estimate the optical properties of tissues and phantom material over a large bandwidth. This knowledge is crucial in order to develop a standardized method, and to provide a-priori information used in the calculation of in vivo chromophores concentration for example. Moreover, the high spectral power combined with the ability to easily and quickly select wavelength make SCL suitable not only for in vivo tissue characterization, and application like breast cancer monitoring, but also to follow fast dynamic physiological processes like cerebral hemodynamic. The large number of wavelength available in the systems based on SCL (several systems reported parallel or quasi-parallel acquisition of up to 16 wavelengths) not only enables to refine the precision of the measurement, but also provides extra information about different processes like metabolism. These sources are also relevant to design instrument aiming at non-contact scanning, which make them also a good candidate for preclinical applications. In conclusion, SCL are a valuable tool at every step of the developmental and translational work, from fundamental characterization to preclinical and clinical use, and we have no doubt that it will still be an important brick to keep driving the innovation in biomedical diffuse optics.

Author Contributions

F.L., L.G. and I.T. conceived and designed the study. F.L. conducted the literature search and drafted the first version of the paper. L.G. created Table 1 and Table 2. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Medical Research Council (MR/S003134/1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Elwell, C.E.; Cooper, C.E. Making light work: Illuminating the future of biomedical optics. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2011, 369, 4358–4379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Maniewsk, R.; Liebert, A.; Kacprzak, M.; Zbieć, A. Selected applications of near infrared optical methods in medical diagnosis. OptoElectron. Rev. 2004, 12, 255–262. [Google Scholar]
  3. Martelli, F.; Binzoni, T.; Pifferi, A.; Spinelli, L.; Farina, A.; Torricelli, A. There’s plenty of light at the bottom: Statistics of photon penetration depth in random media. Sci. Rep. 2016, 6, 27057. [Google Scholar] [CrossRef] [PubMed]
  4. Wolf, M.; Ferrari, M.; Quaresima, V. Progress of near-infrared spectroscopy and topography for brain and muscle clinical applications. J. Biomed. Opt. 2007, 12, 62104. [Google Scholar]
  5. Scholkmann, F.; Kleiser, S.; Metz, A.J.; Zimmermann, R.; Mata Pavia, J.; Wolf, U.; Wolf, M. A review on continuous wave functional near-infrared spectroscopy and imaging instrumentation and methodology. Neuroimage 2014, 85, 6–27. [Google Scholar] [CrossRef] [PubMed]
  6. Yeganeh, H.Z.; Toronov, V.; Elliott, J.T.; Diop, M.; Lee, T.-Y.; St Lawrence, K. Broadband continuous-wave technique to measure baseline values and changes in the tissue chromophore concentrations. Biomed. Opt. Express 2012, 3, 2761–2770. [Google Scholar] [CrossRef] [PubMed]
  7. Bale, G.; Elwell, C.E.; Tachtsidis, I. From Jöbsis to the present day: A review of clinical near-infrared spectroscopy measurements of cerebral cytochrome-c-oxidase. J. Biomed. Opt. 2016, 21, 091307. [Google Scholar] [CrossRef]
  8. Yamada, Y.; Suzuki, H.; Yamashita, Y. Time-Domain Near-Infrared Spectroscopy and Imaging: A Review. Appl. Sci. 2019, 9, 1127. [Google Scholar] [CrossRef] [Green Version]
  9. Grosenick, D.; Wabnitz, H.; Macdonald, R. Diffuse near-infrared imaging of tissue with picosecond time resolution. Biomed. Eng. Biomed. Tech. 2018, 63, 511–518. [Google Scholar] [CrossRef] [PubMed]
  10. Torricelli, A.; Contini, D.; Pifferi, A.; Caffini, M.; Re, R.; Zucchelli, L.; Spinelli, L. Time domain functional NIRS imaging for human brain mapping. Neuroimage 2014, 85, 28–50. [Google Scholar] [CrossRef] [Green Version]
  11. Konugolu Venkata Sekar, S.; Lanka, P.; Farina, A.; Dalla Mora, A.; Andersson-Engels, S.; Taroni, P.; Pifferi, A. Broadband Time Domain Diffuse Optical Reflectance Spectroscopy: A Review of Systems, Methods, and Applications. Appl. Sci. 2019, 9, 5465. [Google Scholar] [CrossRef] [Green Version]
  12. Hoshi, Y.; Yamada, Y. Overview of diffuse optical tomography and its clinical applications. J. Biomed. Opt. 2016, 21, 091312. [Google Scholar] [CrossRef] [Green Version]
  13. Durduran, T.; Choe, R.; Baker, W.B.; Yodh, A.G. Diffuse optics for tissue monitoring and tomography. Rep. Prog. Phys. 2010, 73, 076701. [Google Scholar] [CrossRef] [Green Version]
  14. Labruyère, A.; Tonello, A.; Couderc, V.; Huss, G.; Leproux, P. Compact supercontinuum sources and their biomedical applications. Opt. Fiber Technol. 2012, 18, 375–378. [Google Scholar] [CrossRef]
  15. Alfano, R.R.; Shapiro, S.L. Emission in the Region 4000 to 7000 A Via Four-Photon Coupling in Glass. Phys. Rev. Lett. 1970, 24, 584–587. [Google Scholar] [CrossRef]
  16. Alfano, R.R.; Shapiro, S.L. Observation of Self-Phase Modulation and Small-Scale Filaments in Crystals and Glasses. Phys. Rev. Lett. 1970, 24, 592–594. [Google Scholar] [CrossRef]
  17. Granzow, N. Supercontinuum white light lasers: A review on technology and applications. In Proceedings of the Joint TC1—TC2 International Symposium on Photonics and Education in Measurement Science, Jena, Germany, 17–19 September 2019; Volume 11144, p. 1114408. [Google Scholar]
  18. Bloembergen, N. Nonlinear optics: Past, present, and future. IEEE J. Sel. Top. Quantum Electron. 2000, 6, 876–880. [Google Scholar] [CrossRef]
  19. Agrawal, G.P. Nonlinear fiber optics: Its history and recent progress [Invited]. J. Opt. Soc. Am. B 2011, 28, A1–A10. [Google Scholar] [CrossRef]
  20. Knight, J.C.; Birks, T.A.; Russell, P.S.J.; Atkin, D.M. All-silica single-mode optical fiber with photonic crystal cladding. Opt. Lett. 1996, 21, 1547–1549. [Google Scholar] [CrossRef] [PubMed]
  21. Tu, H.; Boppart, S.A. Coherent fiber supercontinuum for biophotonics. Laser Photonics Rev. 2013, 7, 628–645. [Google Scholar] [CrossRef] [Green Version]
  22. Pifferi, A.; Contini, D.; Mora, A.D.; Farina, A.; Spinelli, L.; Torricelli, A. New frontiers in time-domain diffuse optics, a review. J. Biomed. Opt. 2016, 21, 091310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Konugolu Venkata Sekar, S.; Dalla Mora, A.; Bargigia, I.; Martinenghi, E.; Lindner, C.; Farzam, P.; Pagliazzi, M.; Durduran, T.; Taroni, P.; Pifferi, A.; et al. Broadband (600–1350 nm) Time-Resolved Diffuse Optical Spectrometer for Clinical Use. IEEE J. Sel. Top. Quantum Electron. 2016, 22, 406–414. [Google Scholar] [CrossRef]
  24. Lange, F.; Dunne, L.; Hale, L.; Tachtsidis, I. MAESTROS: A Multiwavelength Time-Domain NIRS System to Monitor Changes in Oxygenation and Oxidation State of Cytochrome-C-Oxidase. IEEE J. Sel. Top. Quantum Electron. 2018, 25. [Google Scholar] [CrossRef] [PubMed]
  25. Giannoni, L.; Lange, F.; Sajic, M.; Smith, K.J.; Tachtsidis, I. A hyperspectral imaging system for mapping haemoglobin and cytochrome-c-oxidase concentration changes in the exposed cerebral cortex. IEEE J. Sel. Top. Quantum Electron. 2021, 27. [Google Scholar] [CrossRef] [PubMed]
  26. Jiang, J.; Mata, A.D.C.; Lindner, S.; Charbon, E.; Wolf, M.; Kalyanov, A. Dynamic time domain near-infrared optical tomography based on a SPAD camera. Biomed. Opt. Express 2020, 11, 5470. [Google Scholar] [CrossRef]
  27. Jiang, J.; Mata, A.D.C.; Lindner, S.; Zhang, C.; Charbon, E.; Wolf, M.; Kalyanov, A. Image reconstruction for novel time domain near infrared optical tomography: Towards clinical applications. Biomed. Opt. Express 2020, 11, 4723. [Google Scholar] [CrossRef] [PubMed]
  28. Cooper, R.J.; Magee, E.; Everdell, N.; Magazov, S.; Varela, M.; Airantzis, D.; Gibson, A.P.; Hebden, J.C. MONSTIR II: A 32-channel, multispectral, time-resolved optical tomography system for neonatal brain imaging. Rev. Sci. Instrum. 2014, 85. [Google Scholar] [CrossRef] [PubMed]
  29. Andersson-Engels, S.; Berg, R.; Persson, A.; Svanberg, S. Multispectral tissue characterization with time-resolved detection of diffusely scattered white light. Opt. Lett. 1993, 18, 1697. [Google Scholar] [CrossRef]
  30. af Klinteberg, C.; Berg, R.; Lindquist, C.; Andersson-Engels, S.; Svanberg, S. Diffusely scattered femtosecond white-light examination of breast tissue in vitro and in vivo. Photon Propag. Tissues 1995, 2626, 149–157. [Google Scholar]
  31. Jarlman, O.; Berg, R.; Andersson-Engels, S.; Svanberg, S.; Pettersson, H. Time-resolved white light transillumination for optical imaging. Acta Radiol. 1997, 38, 185–189. [Google Scholar] [CrossRef]
  32. Abrahamsson, C.; Svensson, T.; Svanberg, S.; Andersson-Engels, S.; Johansson, J.; Folestad, S. Time and wavelength resolved spectroscopy of turbid media using light continuum generated in a crystal fiber. Opt. Express 2004, 12, 4103–4112. [Google Scholar] [CrossRef] [Green Version]
  33. Pifferi, A.; Torricelli, A.; Taroni, P.; Comelli, D.; Bassi, A.; Cubeddu, R. Fully automated time domain spectrometer for the absorption and scattering characterization of diffusive media. Rev. Sci. Instrum. 2007, 78, 053103. [Google Scholar] [CrossRef] [PubMed]
  34. Taroni, P.; Pifferi, A.; Torricelli, A.; Comelli, D.; Cubeddu, R. In vivo absorption and scattering spectroscopy of biological tissues. Photochem. Photobiol. Sci. 2003, 2, 124–129. [Google Scholar] [CrossRef]
  35. Pifferi, A.; Torricelli, A.; Taroni, P.; Bassi, A.; Chikoidze, E.; Giambattistelli, E.; Cubeddu, R. Optical biopsy of bone tissue: A step toward the diagnosis of bone pathologies. J. Biomed. Opt. 2004, 9, 474. [Google Scholar] [CrossRef] [PubMed]
  36. Cubeddu, R.; Pifferi, A.; Taroni, P.; Torricelli, A.; Valentini, G. Noninvasive absorption and scattering spectroscopy of bulk diffusive media: An application to the optical characterization of human breast. Appl. Phys. Lett. 1999, 74, 874–876. [Google Scholar] [CrossRef]
  37. Taroni, P.; Pifferi, A.; Cubeddu, R.; Torricelli, A.; Comelli, D. Absorption of collagen: Effects on the estimate of breast composition and related diagnostic implications. J. Biomed. Opt. 2007, 12, 014021. [Google Scholar] [CrossRef]
  38. Bassi, A.; Farina, A.; D’Andrea, C.; Pifferi, A.; Valentini, G.; Cubeddu, R. Portable, large-bandwidth time-resolved system for diffuse optical spectroscopy. Opt. Express 2007, 15, 14482. [Google Scholar] [CrossRef]
  39. Taroni, P.; Bassi, A.; Comelli, D.; Farina, A.; Cubeddu, R.; Pifferi, A. Diffuse optical spectroscopy of breast tissue extended to 1100 nm. J. Biomed. Opt. 2009, 14, 054030. [Google Scholar] [CrossRef] [PubMed]
  40. Farina, A.; Bassi, A.; Pifferi, A.; Taroni, P.; Comelli, D.; Spinelli, L.; Cubeddu, R. Bandpass Effects in Time-Resolved Diffuse Spectroscopy. Appl. Spectrosc. 2009, 63, 48–56. [Google Scholar] [CrossRef] [Green Version]
  41. Farina, A.; Bargigia, I.; Taroni, P.; Pifferi, A. Note: Comparison between a prism-based and an acousto-optic tunable filter-based spectrometer for diffusive media. Rev. Sci. Instrum. 2013, 84. [Google Scholar] [CrossRef]
  42. Svensson, T.; Alerstam, E.; Khoptyar, D.; Johansson, J.; Folestad, S.; Andersson-Engels, S. Near-infrared photon time-of-flight spectroscopy of turbid materials up to 1400 nm. Rev. Sci. Instrum. 2009, 80, 45–48. [Google Scholar] [CrossRef] [Green Version]
  43. Bargigia, I.; Tosi, A.; Farina, A.; Bassi, A.; Taroni, P.; Bahgat Shehata, A.; Della Frera, A.; Dalla Mora, A.; Zappa, F.; Cubeddu, R.; et al. Optical Spectroscopy up to 1700 nm: A Time-Resolved Approach Combined with an InGaAs/InP Single-Photon Avalanche Diode. In Proceedings of the Biomedical Optics and 3-D Imaging 2012, Miami, FL, USA, 28 April–2 May 2012; p. JM3A.16. [Google Scholar]
  44. Bargigia, I.; Tosi, A.; Bahgat Shehata, A.; Della Frera, A.; Farina, A.; Bassi, A.; Taroni, P.; Dalla Mora, A.; Zappa, F.; Cubeddu, R.; et al. Time-resolved diffuse optical spectroscopy up to 1700 nm by means of a time-gated InGaAs/InP single-photon avalanche diode. Appl. Spectrosc. 2012, 66, 944–950. [Google Scholar] [CrossRef]
  45. Bargigia, I.; Tosi, A.; Bahgat Shehata, A.; Della Frera, A.; Farina, A.; Bassi, A.; Taroni, P.; Dalla Mora, A.; Zappa, F.; Cubeddu, R.; et al. In-vivo optical spectroscopy in the time-domain beyond 1100 nm. In Proceedings of the Optics InfoBase Conference Papers, Munich, Germany, 12–16 May 2013; Taroni, P., Dehghani, H., Eds.; Optical Society of America: Washington, DC, USA, 2013; p. 879902. [Google Scholar]
  46. Sekar, S.K.V.; Bargigia, I.; Mora, A.D.; Taroni, P.; Ruggeri, A.; Tosi, A.; Pifferi, A.; Farina, A. Diffuse optical characterization of collagen absorption from 500 to 1700 nm. J. Biomed. Opt. 2017, 22, 015006. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Konugolu Venkata Sekar, S.; Farina, A.; Martinenghi, E.; Dalla Mora, A.; Taroni, P.; Pifferi, A.; Negredo, E.; Puig, J.; Escrig, R.; Rosales, Q.; et al. Time-resolved diffused optical characterization of key tissue constituents of human bony prominence locations. In Proceedings of the European Conference on Biomedical Optics, Munich, Germany, 21–25 June 2015; Volume 9538, pp. 95380X–95380X–5. [Google Scholar]
  48. Konugolu Venkata Sekar, S.; Pagliazzi, M.; Negredo, E.; Martelli, F.; Farina, A.; Dalla Mora, A.; Lindner, C.; Farzam, P.; Pérez-Álvarez, N.; Puig, J.; et al. In Vivo, non-invasive characterization of human bone by hybrid broadband (600–1200 nm) diffuse optical and correlation spectroscopies. PLoS ONE 2016, 11, e0168426. [Google Scholar] [CrossRef] [Green Version]
  49. Sekar, S.K.V.; Mora, A.D.; Martinenghi, E.; Taroni, P.; Pifferi, A.; Farina, A.; Puig, J.; Negredo, E.; Lindner, C.; Pagliazzi, M.; et al. In vivo Time domain Broadband (600–1200 nm) Diffuse Optical Characterization of Human Bone. In Proceedings of the Optical Tomography and Spectroscopy 2016, Fort Lauderdale, FL, USA, 25–28 April 2016; p. JTu3A.32. [Google Scholar]
  50. Konugolu Venkata Sekar, S.; Farina, A.; Dalla Mora, A.; Lindner, C.; Pagliazzi, M.; Mora, M.; Aranda, G.; Dehghani, H.; Durduran, T.; Taroni, P.; et al. Broadband (550–1350 nm) diffuse optical characterization of thyroid chromophores. Sci. Rep. 2018, 8, 1–8. [Google Scholar] [CrossRef] [Green Version]
  51. Lanka, P.; Segala, A.; Farina, A.; Konugolu Venkata Sekar, S.; Nisoli, E.; Valerio, A.; Taroni, P.; Cubeddu, R.; Pifferi, A. Non-invasive investigation of adipose tissue by time domain diffuse optical spectroscopy. Biomed. Opt. Express 2020, 11, 2779. [Google Scholar] [CrossRef] [PubMed]
  52. Taroni, P.; Bassi, A.; Farina, A.; Cubeddu, R.; Pifferi, A. Role of Collagen Scattering for in vivo Tissue Characterization. In Proceedings of the Biomedical Optics 2010, Miami, FL, USA, 11–14 April 2010; p. BTuD107. [Google Scholar]
  53. Konugolu Venkata Sekar, S.; Beh, J.S.; Farina, A.; Dalla Mora, A.; Pifferi, A.; Taroni, P. Broadband diffuse optical characterization of elastin for biomedical applications. Biophys. Chem. 2017, 229, 130–134. [Google Scholar] [CrossRef] [PubMed]
  54. Ferocino, E.; Di Sciacca, G.; Di Sieno, L.; Dalla Mora, A.; Pifferi, A.; Arridge, S.R.; Martelli, F.; Taroni, P.; Farina, A. Multi-wavelength time domain diffuse optical tomography for breast cancer: Initial results on silicone phantoms. In Proceedings of the SPIE BiOS, San Francisco, CA, USA, 2–7 February 2019; p. 59. [Google Scholar]
  55. Taroni, P.; Pifferi, A.; Quarto, G.; Farina, A.; Ieva, F.; Paganoni, A.M.; Abbate, F.; Cassano, E.; Cubeddu, R. Time domain diffuse optical spectroscopy: In vivo quantification of collagen in breast tissue. In Proceedings of the Optical Methods for Inspection, Characterization, and Imaging of Biomaterials II, Munich, Germany, 21–25 June 2015; Volume 9529, p. 952910. [Google Scholar]
  56. Lanka, P.; Joseph, F.K.; Kruit, H.; Konugolu, S.; Sekar, V.; Farina, A.; Cubeddu, R.; Manohar, S.; Pifferi, A. Time domain diffuse optical spectroscopy for the monitoring of thermal treatment in biological tissue. In Proceedings of the Optical Tomography and Spectroscopy 2020, Washington, DC, USA, 20–23 April 2020; pp. 1–2. [Google Scholar]
  57. Kalloor Joseph, F.; Lanka, P.; Kruit, H.; Konugolu Venkata Sekar, S.; Farina, A.; Cubeddu, R.; Manohar, S.; Pifferi, A. Key features in the optical properties of tissue during and after radiofrequency ablation. In Proceedings of the SPIE BiOS, San Francisco, CA, USA, 1–6 February 2020; p. 16. [Google Scholar]
  58. Bassi, A.; Spinelli, L.; D’Andrea, C.; Giusto, A.; Swartling, J.; Pifferi, A.; Torricelli, A.; Cubeddu, R. Feasibility of white-light time-resolved optical mammography. J. Biomed. Opt. 2006, 11, 054035. [Google Scholar] [CrossRef]
  59. Ferrari, M.; Mottola, L.; Quaresima, V. Principles, techniques, and limitations of near infrared spectroscopy. Can. J. Appl. Physiol. 2004, 29, 463–487. [Google Scholar] [CrossRef] [Green Version]
  60. Bassi, A.; Swartling, J.; D’Andrea, C.; Pifferi, A.; Torricelli, A.; Cubeddu, R. Time-resolved spectrophotometer for turbid media based on supercontinuum generation in a photonic crystal fiber. Opt. Lett. 2004, 29, 2405. [Google Scholar] [CrossRef]
  61. Swartling, J.; Bassi, A.; D’Andrea, C.; Pifferi, A.; Torricelli, A.; Cubeddu, R. Dynamic time-resolved diffuse spectroscopy based on supercontinuum light pulses. Appl. Opt. 2005, 44, 4684. [Google Scholar] [CrossRef] [PubMed]
  62. Gerega, A.; Milej, D.; Weigl, W.; Kacprzak, M.; Liebert, A. Multiwavelength time-resolved near-infrared spectroscopy of the adult head: Assessment of intracerebral and extracerebral absorption changes. Biomed. Opt. Express 2018, 9, 2974. [Google Scholar] [CrossRef]
  63. Gerega, A.; Milej, D.; Weigl, W.; Zolek, N.; Sawosz, P.; Maniewski, R.; Liebert, A. Multi-wavelength time-resolved measurements of diffuse reflectance: Phantom study with dynamic inflow of ICG. In Proceedings of the Biomedical Optics 2012, Miami, FL, USA, 28 April–2 May 2012; p. JM3A.31. [Google Scholar]
  64. Sudakou, A.; Lange, F.; Isler, H.; Gerega, A.; Ostojic, D.; Sawosz, P.; Tachtsidis, I.; Wolf, M.; Liebert, A. Multi-wavelength time-resolved NIRS measurements for estimation of absolute concentration of chromophores: Blood phantom study. In Proceedings of the Diffuse Optical Spectroscopy and Imaging VII, Munich, Germany, 23–25 June 2019; Dehghani, H., Wabnitz, H., Eds.; SPIE: Washington, DC, USA, 2019; Volume 1107422, p. 72. [Google Scholar]
  65. He, L.; Baker, W.B.; Busch, D.R.; Jiang, J.Y.; Lawrence, K.S.; Kofke, W.A.; Yodh, A.G.; He, L.; Baker, W.B.; Milej, D.; et al. Noninvasive continuous optical monitoring of absolute cerebral blood flow in critically ill adults. Neurophotonics 2018, 5, 1. [Google Scholar]
  66. Durduran, T.; Yodh, A.G. Diffuse correlation spectroscopy for non-invasive, micro-vascular cerebral blood flow measurement. Neuroimage 2014, 85, 5163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Selb, J.; Zimmermann, B.B.; Martino, M.; Ogden, T.; Boas, D.A. Functional brain imaging with a supercontinuum time-domain NIRS system. In Proceedings of the Optical Tomography and Spectroscopy of Tissue X, San Francisco, CA, USA, 5–7 February 2013; Volume 8578, p. 857807. [Google Scholar]
  68. Selb, J.; Joseph, D.K.; Boas, D. Time-gated optical system for depth-resolved functional brain imaging. J. Biomed. Opt. 2006, 11, 44008. [Google Scholar] [CrossRef]
  69. Lange, F.; Peyrin, F.; Montcel, B. Broadband time-resolved multi-channel functional near-infrared spectroscopy system to monitor in vivo physiological changes of human brain activity. Appl. Opt. 2018, 57, 6417. [Google Scholar] [CrossRef]
  70. Lange, F.; Peyrin, F.; Montcel, B. A hyperspectral time resolved DOT system to monitor physiological changes of the human brain activity. In Proceedings of the Advanced Microscopy Techniques IV and Neurophotonics II, Munich, Germany, 21–25 June 2015; OSA: Washington, DC, USA, 2015; p. 95360R. [Google Scholar]
  71. Lange, F.; Dunne, L.; Tachtsidis, I. Evaluation of Haemoglobin and Cytochrome Responses During Forearm Ischaemia Using Multi-wavelength Time Domain NIRS. In Oxygen Transport to Tissue XXXIX; Advances in Experimental Medicine and Biology; Halpern, H.J., LaManna, J.C., Harrison, D.K., Epel, B., Eds.; Springer International Publishing: Cham, Switzerland, 2017; Volume 977, pp. 67–72. ISBN 978-3-319-55229-3. [Google Scholar]
  72. Lange, F.; Tachtsidis, I. Short and mid-term reproducibility analysis of cerebral tissue saturation measured by time domain-NIRS. In Proceedings of the European Conference on Biomedical Optics, Munich, Germany, 23–25 June 2019; Volume 11074. [Google Scholar]
  73. Mazurenka, M.; Jelzow, A.; Wabnitz, H.; Contini, D.; Spinelli, L.; Pifferi, A.; Cubeddu, R.; Mora, A.D.; Tosi, A.; Zappa, F.; et al. Non-contact time-resolved diffuse reflectance imaging at null source-detector separation. Opt. Express 2011, 20, 283. [Google Scholar] [CrossRef]
  74. Wabnitz, H.; Mazurenka, M.; Di Sieno, L.; Contini, D.; Dalla Mora, A.; Farina, A.; Hoshi, Y.; Kirilina, E.; Macdonald, R.; Pifferi, A. Non-contact time-domain imaging of functional brain activation and heterogeneity of superficial signals. In Proceedings of the Diffuse Optical Spectroscopy and Imaging VI, Munich, Germany, 25–29 June 2017; Dehghani, H., Wabnitz, H., Eds.; SPIE: Washington, DC, USA, 2017; Volume 10412, p. 104120J. [Google Scholar]
  75. Mazurenka, M.; Di Sieno, L.; Boso, G.; Contini, D.; Pifferi, A.; Mora, A.D.; Tosi, A.; Wabnitz, H.; Macdonald, R. Non-contact in vivo diffuse optical imaging using a time-gated scanning system. Biomed. Opt. Express 2013, 4, 2257. [Google Scholar] [CrossRef]
  76. Di Sieno, L.; Wabnitz, H.; Pifferi, A.; Mazurenka, M.; Hoshi, Y.; Dalla Mora, A.; Contini, D.; Boso, G.; Becker, W.; Martelli, F.; et al. Characterization of a time-resolved non-contact scanning diffuse optical imaging system exploiting fast-gated single-photon avalanche diode detection. Rev. Sci. Instrum. 2016, 87. [Google Scholar] [CrossRef] [Green Version]
  77. Torricelli, A.; Pifferi, A.; Spinelli, L.; Cubeddu, R.; Martelli, F.; Del Bianco, S.; Zaccanti, G. Time-Resolved Reflectance at Null Source-Detector Separation: Improving Contrast and Resolution in Diffuse Optical Imaging. Phys. Rev. Lett. 2005, 95, 078101. [Google Scholar] [CrossRef]
  78. Dalla Mora, A.; Tosi, A.; Zappa, F.; Cova, S.; Contini, D.; Pifferi, A.; Spinelli, L.; Torricelli, A.; Cubeddu, R. Fast-Gated Single-Photon Avalanche Diode for Wide Dynamic Range Near Infrared Spectroscopy. IEEE J. Sel. Top. Quantum Electron. 2010, 16, 1023–1030. [Google Scholar] [CrossRef]
  79. Contini, D.; Dalla Mora, A.; Spinelli, L.; Farina, A.; Torricelli, A.; Cubeddu, R.; Martelli, F.; Zaccanti, G.; Tosi, A.; Boso, G.; et al. Effects of time-gated detection in diffuse optical imaging at short source-detector separation. J. Phys. D Appl. Phys. 2015, 48, 045401. [Google Scholar] [CrossRef] [Green Version]
  80. Mottin, S.; Montcel, B.; de Chatellus, H.G.; Ramstein, S.; Vignal, C.; Mathevon, N. Functional White-Laser Imaging to Study Brain Oxygen Uncoupling/Recoupling in Songbirds. J. Cereb. Blood Flow Metab. 2011, 31, 393–400. [Google Scholar] [CrossRef] [Green Version]
  81. Ramstein, S.; Vignal, C.; Mathevon, N.; Mottin, S. In vivo and noninvasive measurement of a songbird head’s optical properties. Appl. Opt. 2005, 44, 6197. [Google Scholar] [CrossRef] [Green Version]
  82. Torabzadeh, M.; Stockton, P.; Kennedy, G.T.; Saager, R.B.; Durkin, A.J.; Bartels, R.A.; Tromberg, B.J. Hyperspectral imaging in the spatial frequency domain with a supercontinuum source. J. Biomed. Opt. 2019, 24, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Farina, A.; Lepore, M.; Di Sieno, L.; Dalla Mora, A.; Ducros, N.; Pifferi, A.; Valentini, G.; Arridge, S.; D’Andrea, C. Diffuse optical tomography by using time-resolved single pixel camera. In Proceedings of the SPIE BiOS, San Francisco, CA, USA, 7–12 February 2015; Tromberg, B.J., Yodh, A.G., Sevick-Muraca, E.M., Alfano, R.R., Eds.; International Society for Optics and Photonics: Bellingham, WA, USA, 2015; p. 93191K. [Google Scholar]
  84. Pian, Q.; Yao, R.; Intes, X. Hyperspectral wide-field time domain single-pixel diffuse optical tomography platform. Biomed. Opt. Express 2018, 9, 6258. [Google Scholar] [CrossRef]
  85. Valim, N.; Brock, J.; Niedre, M. Experimental measurement of time-dependent photon scatter for diffuse optical tomography. J. Biomed. Opt. 2010, 15, 065006. [Google Scholar] [CrossRef] [Green Version]
  86. Zouaoui, J.; Di Sieno, L.; Hervé, L.; Pifferi, A.; Farina, A.; Mora, A.D.; Derouard, J.; Dinten, J.-M. Chromophore decomposition in multispectral time-resolved diffuse optical tomography. Biomed. Opt. Express 2017, 8, 4772. [Google Scholar] [CrossRef] [Green Version]
  87. Di Sieno, L.; Bettega, G.; Berger, M.; Hamou, C.; Aribert, M.; Mora, A.D.; Puszka, A.; Grateau, H.; Contini, D.; Hervé, L.; et al. Toward noninvasive assessment of flap viability with time-resolved diffuse optical tomography: A preclinical test on rats. J. Biomed. Opt. 2016, 21, 025004. [Google Scholar] [CrossRef] [Green Version]
  88. Dempsey, L.A.; Cooper, R.J.; Powell, S.; Edwards, A.; Lee, C.-W.; Brigadoi, S.; Everdell, N.; Arridge, S.; Gibson, A.P.; Austin, T.; et al. Whole-head functional brain imaging of neonates at cot-side using time-resolved diffuse optical tomography. In Proceedings of the Diffuse Optical Imaging V, Munich, Germany, 21–25 June 2015; Volume 9538, pp. 1–10. [Google Scholar]
  89. Dempsey, L.A. Development and Application of Diffuse Optical Tomography Systems for Diagnosis and Assessment of Perinatal Brain Injury. Ph.D. Thesis, University College London, London, UK, 2018. [Google Scholar]
  90. Jiang, J.; Ahnen, L.; Kalyanov, A.; Lindner, S.; Wolf, M.; Majos, S.S. A New Method Based on Graphics Processing Units for Fast Near-Infrared Optical Tomography. Adv. Exp. Med. Biol. 2017, 977, 191–197. [Google Scholar]
  91. Arnesano, C.; Santoro, Y.; Gratton, E. Digital parallel frequency-domain spectroscopy for tissue imaging. J. Biomed. Opt. 2012, 17, 0960141. [Google Scholar] [CrossRef] [PubMed]
  92. Pifferi, A.; Torricelli, A.; Bassi, A.; Taroni, P.; Cubeddu, R.; Wabnitz, H.; Grosenick, D.; Möller, M.; Macdonald, R.; Swartling, J.; et al. Performance assessment of photon migration instruments: The MEDPHOT protocol. Appl. Opt. 2005, 44, 2104–2114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Wabnitz, H.; Jelzow, A.; Mazurenka, M.; Steinkellner, O.; Macdonald, R.; Milej, D.; Zolek, N.; Kacprzak, M.; Sawosz, P.; Maniewski, R.; et al. Performance assessment of time-domain optical brain imagers, part 1: Basic instrumental performance protocol. J. Biomed. Opt. 2014, 19, 86010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Wabnitz, H.; Jelzow, A.; Mazurenka, M.; Steinkellner, O.; Macdonald, R.; Milej, D.; Zolek, N.; Kacprzak, M.; Sawosz, P.; Maniewski, R.; et al. Performance assessment of time-domain optical brain imagers, part 2: nEUROPt protocol. J. Biomed. Opt. 2014, 19, 086012. [Google Scholar] [CrossRef]
  95. Quarto, G.; Pifferi, A.; Bargigia, I.; Farina, A.; Cubeddu, R.; Taroni, P. Recipes to make organic phantoms for diffusive optical spectroscopy. Appl. Opt. 2013, 52, 2494–2502. [Google Scholar] [CrossRef]
  96. Wang, L.; Sharma, S.; Aernouts, B.; Ramon, H.; Saeys, W. Supercontinuum laser based double-integrating-sphere system for measuring optical properties of highly dense turbid media in the 1300–2350 nm region with high sensitivity. In Proceedings of the SPIE Photonics Europe, Brussels, Belgium, 16–19 April 2012; Volume 8427, p. 84273B. [Google Scholar]
  97. Spinelli, L.; Botwicz, M.; Zolek, N.; Kacprzak, M.; Milej, D.; Liebert, A.; Weigel, U.; Durduran, T.; Foschum, F.; Kienle, A.; et al. Inter-Laboratory Comparison of Optical Properties Performed on Intralipid and India Ink. In Proceedings of the Biomedical Optics and 3-D Imaging, Miami, FL, USA, 28 April–2 May 2012; OSA: Washington, DC, USA, 2012; p. BW1A.6. [Google Scholar]
  98. Aernouts, B.; Zamora-Rojas, E.; Van Beers, R.; Watté, R.; Wang, L.; Tsuta, M.; Lammertyn, J.; Saeys, W. Supercontinuum laser based optical characterization of Intralipid® phantoms in the 500–2250 nm range. Opt. Express 2013, 21, 32450. [Google Scholar] [CrossRef]
  99. Aernouts, B.; Van Beers, R.; Watté, R.; Lammertyn, J.; Saeys, W. Dependent scattering in Intralipid® phantoms in the 600–1850 nm range. Opt. Express 2014, 22, 6086. [Google Scholar] [CrossRef]
  100. Spinelli, L.; Botwicz, M.; Zolek, N.; Kacprzak, M.; Milej, D.; Sawosz, P.; Liebert, A.; Weigel, U.; Durduran, T.; Foschum, F.; et al. Determination of reference values for optical properties of liquid phantoms based on Intralipid and India ink. Biomed. Opt. Express 2014, 5, 2037. [Google Scholar] [CrossRef]
  101. Bouchard, J.-P.; Veilleux, I.; Jedidi, R.; Noiseux, I.; Fortin, M.; Mermut, O. Reference optical phantoms for diffuse optical spectroscopy. Part 1—Error analysis of a time resolved transmittance characterization method. Opt. Express 2010, 18, 11495–11507. [Google Scholar] [CrossRef]
  102. Martelli, F.; Di Ninni, P.; Zaccanti, G.; Contini, D.; Spinelli, L.; Torricelli, A.; Cubeddu, R.; Wabnitz, H.; Mazurenka, M.; Macdonald, R.; et al. Phantoms for diffuse optical imaging based on totally absorbing objects, part 2: Experimental implementation. J. Biomed. Opt. 2014, 19, 076011. [Google Scholar] [CrossRef] [Green Version]
  103. Pifferi, A.; Torricelli, A.; Cubeddu, R.; Quarto, G.; Re, R.; Sekar, S.K.V.; Spinelli, L.; Farina, A.; Martelli, F.; Wabnitz, H. Mechanically switchable solid inhomogeneous phantom for performance tests in diffuse imaging and spectroscopy. J. Biomed. Opt. 2015, 20, 121304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Wabnitz, H.; Taubert, D.R.; Funane, T.; Kiguchi, M.; Eda, H.; Pifferi, A.; Torricelli, A.; Macdonald, R. Characterization of homogeneous tissue phantoms for performance tests in diffuse optics. In Proceedings of the SPIE BiOS, San Francisco, CA, USA, 13–18 February 2016; Volume 9700, p. 970004. [Google Scholar]
  105. Xu, R.X.; Allen, D.W.; Huang, J.; Gnyawali, S.; Melvin, J.; Elgharably, H.; Gordillo, G.; Huang, K.; Bergdall, V.; Litorja, M.; et al. Developing digital tissue phantoms for hyperspectral imaging of ischemic wounds. Biomed. Opt. Express 2012, 3, 1433. [Google Scholar] [CrossRef] [Green Version]
  106. Wabnitz, H.; Hwang, J.; Yang, L.; Macdonald, R. Digital phantom for time-domain near-infrared spectroscopy of tissue: Concept and proof-of-principle experiments. In Proceedings of the SPIE BiOS, San Francisco, CA, USA, 2–7 February 2019; p. 20. [Google Scholar]
  107. Spinelli, L.; Rizzolo, A.; Vanoli, M.; Grassi, M.; Zerbini, P.E.; Pimentel, R.; Torricelli, A. Optical properties of pulp and skin in Brazilian mangoes in the 540–900 nm spectral range: Implication for non-destructive maturity assessment by time-resolved reflectance spectroscopy. In Proceedings of the International Conference of Agricultural Engineering, CIGR-AgEng2012, Valencia, Spain, 8–12 July 2012; p. C–2096. [Google Scholar]
  108. Amendola, C.; Pirovano, I.; Lacerenza, M.; Contini, D.; Spinelli, L.; Cubeddu, R.; Torricelli, A.; Re, R. Use of 3D printed PLA for diffuse optics. In Proceedings of the Biophotonics Congress: Biomedical Optics 2020 (Translational, Microscopy, OCT, OTS, BRAIN), Washington, DC, USA, 20–23 April 2020; pp. 32–33. [Google Scholar]
  109. Dalla Mora, A.; Martinenghi, E.; Contini, D.; Tosi, A.; Boso, G.; Durduran, T.; Arridge, S.; Martelli, F.; Farina, A.; Torricelli, A.; et al. Fast silicon photomultiplier improves signal harvesting and reduces complexity in time-domain diffuse optics. Opt. Express 2015, 23, 13937–13946. [Google Scholar] [CrossRef] [Green Version]
  110. Martinenghi, E.; Dalla Mora, A.; Contini, D.; Farina, A.; Villa, F.; Torricelli, A.; Pifferi, A. Spectrally Resolved Single-Photon Timing of Silicon Photomultipliers for Time-Domain Diffuse Spectroscopy. IEEE Photonics J. 2015, 7, 1–12. [Google Scholar] [CrossRef] [Green Version]
  111. Martinenghi, E.; Di Sieno, L.; Contini, D.; Sanzaro, M.; Pifferi, A.; Dalla Mora, A. Time-resolved single-photon detection module based on silicon photomultiplier: A novel building block for time-correlated measurement systems. Rev. Sci. Instrum. 2016, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Di Sieno, L.; Boetti, N.G.; Mora, A.D.; Pugliese, D.; Farina, A.; Konugolu Venkata Sekar, S.; Ceci-Ginistrelli, E.; Janner, D.; Pifferi, A.; Milanese, D. Towards the use of bioresorbable fibers in time-domain diffuse optics. J. Biophotonics 2017, 11, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Mora, A.D.; Di Sieno, L.; Venkata Sekar, S.K.; Farina, A.; Contini, D.; Boetti, N.G.; Milanese, D.; Nissinen, J.; Pifferi, A. Novel technologies for time-domain diffuse optics: Miniaturized wearable devices and bioresorbable optical fibers. In Proceedings of the Biophotonics Congress: Biomedical Optics Congress 2018 (Microscopy/Translational/Brain/OTS), Hollywood, FL, USA, 3–6 April 2018; p. F90-O. [Google Scholar]
  114. D’Andrea, C.; Spinelli, L.; Bassi, A.; Giusto, A.; Contini, D.; Swartling, J.; Torricelli, A.; Cubeddu, R. Time-resolved spectrally constrained method for the quantification of chromophore concentrations and scattering parameters in diffusing media. Opt. Express 2006, 14, 1888. [Google Scholar] [CrossRef] [PubMed]
  115. Giusto, A.; D’Andrea, C.; Spinelli, L.; Contini, D.; Torricelli, A.; Martelli, F.; Zaccanti, G.; Cubeddu, R. Monitoring absorption changes in a layered diffusive medium by white-light time-resolved reflectance spectroscopy. IEEE Trans. Instrum. Meas. 2010, 59, 1925–1932. [Google Scholar] [CrossRef]
  116. Pifferi, A.; Bassi, A.; Spinelli, L.; Cubeddu, R.; Taroni, P. Time-Resolved Broadband Diffuse Spectroscopy Using a Differential Absorption Approach. In Proceedings of the Biomedical Optics and 3-D Imaging, Miami, FL, USA, 11–14 April 2010; p. BSuD45; Volume 64, p. BSuD45. [Google Scholar]
  117. Yang, L.; Wabnitz, H.; Gladytz, T.; Macdonald, R.; Grosenick, D. Spatially-enhanced time-domain NIRS for accurate determination of tissue optical properties. Opt. Express 2019, 27, 26415. [Google Scholar] [CrossRef] [PubMed]
  118. Yang, L.; Lanka, P.; Wabnitz, H.; Cubeddu, R.; Gladytz, T.; Konugolu Venkata Sekar, S.; Grosenick, D.; Pifferi, A.; Macdonald, R. Spatially-enhanced time-domain NIRS for determination of optical properties in layered structures. In Proceedings of the European Conference on Biomedical Optics, Munich, Germany, 23–25 June 2019; p. 9. [Google Scholar]
  119. Yang, L.; Wabnitz, H.; Gladytz, T.; Sudakou, A.; Macdonald, R.; Grosenick, D. Space-enhanced time-domain diffuse optics for determination of tissue optical properties in two-layered structures. Biomed. Opt. Express 2020, 11, 6570. [Google Scholar] [CrossRef]
  120. Wojtkiewicz, S.; Gerega, A.; Zanoletti, M.; Sudakou, A.; Contini, D.; Liebert, A.; Durduran, T.; Dehghani, H. Self-calibrating time-resolved near infrared spectroscopy. Biomed. Opt. Express 2019, 10, 2657. [Google Scholar] [CrossRef]
  121. Kovacsova, Z.; Bale, G.; Mitra, S.; Lange, F.; Tachtsidis, I. Absolute quantification of cerebral tissue oxygen saturation with multidistance broadband NIRS in newborn brain. Biomed. Opt. Express 2021, 12, 907. [Google Scholar] [CrossRef]
  122. Wabnitz, H.; Contini, D.; Spinelli, L.; Torricelli, A.; Liebert, A. Depth-selective analysis in time-domain optical brain imaging: Moments vs. time windows. Biomed. Opt. Express 2020, 11, 4224–4243. [Google Scholar] [CrossRef] [PubMed]
  123. Sudakou, A.; Yang, L.; Wabnitz, H.; Wojtkiewicz, S.; Liebert, A. Performance of measurands in time-domain optical brain imaging: Depth selectivity versus contrast-to-noise ratio. Biomed. Opt. Express 2020, 11, 4348–4365. [Google Scholar] [CrossRef]
  124. Lanka, P.; Yang, L.; Orive-Miguel, D.; Veesa, J.D.; Tagliabue, S.; Sudakou, A.; Samaei, S.; Forcione, M.; Kovacsova, Z.; Behera, A.; et al. The BITMAP exercise: A multi-laboratory performance assessment campaign of diffuse optical instrumentation. In Proceedings of the Diffuse Optical Spectroscopy and Imaging VII, Munich, Germany, 23–25 June 2019; Dehghani, H., Wabnitz, H., Eds.; SPIE: Washington, DC, USA; p. 44. [Google Scholar]
  125. Orive-Miguel, D.; Lanka, P.; Yang, L.; Tagliabue, S.; Sudakou, A.; Samaei, S.; Veesa, J.D.; Forcione, M.; Kovacsova, Z.; Behera, A.; et al. The BitMap dataset: An open dataset on performance assessment of diffuse optics instruments. In Proceedings of the European Conference on Biomedical Optics 2019, Munich, Germany, 23–25 June 2019; p. 45. [Google Scholar]
  126. Farina, A.; Pifferi, A.; Torricelli, A.; Bargigia, I.; Spinelli, L.; Cubeddu, R.; Foschum, F.; Jäger, M.; Simon, E.; Fugger, O.; et al. Multi-center study of the optical properties of the human head. In Proceedings of the Biomedical Optics 2014, Miami, FL, USA, 26–30 April 2014; Volume 6, p. BS3A.9. [Google Scholar]
  127. Sordillo, L.A.; Lindwasser, L.; Budansky, Y.; Leproux, P.; Alfano, R.R. Near-infrared supercontinuum laser beam source in the second and third near-infrared optical windows used to image more deeply through thick tissue as compared with images from a lamp source. J. Biomed. Opt. 2015, 20, 030501. [Google Scholar] [CrossRef] [PubMed]
  128. Yadav, J.; Rani, A.; Singh, V.; Murari, B.M. Prospects and limitations of non-invasive blood glucose monitoring using near-infrared spectroscopy. Biomed. Signal. Process. Control 2015, 18, 214–227. [Google Scholar] [CrossRef]
  129. Rostami, E. Glucose and the injured brain-monitored in the neurointensive care unit. Front. Neurol. 2014, 5. [Google Scholar] [CrossRef] [Green Version]
  130. Pinchefsky, E.F.; Hahn, C.D.; Kamino, D.; Chau, V.; Brant, R.; Moore, A.M.; Tam, E.W.Y. Hyperglycemia and Glucose Variability Are Associated with Worse Brain Function and Seizures in Neonatal Encephalopathy: A Prospective Cohort Study. J. Pediatr. 2019, 209, 23–32. [Google Scholar] [CrossRef] [PubMed]
  131. Liu, J.; Zhu, C.; Jiang, J.; Xu, K. Scattering-independent glucose absorption measurement using a spectrally resolved reflectance setup with specialized variable source-detector separations. Biomed. Opt. Express 2018, 9, 5903. [Google Scholar] [CrossRef]
  132. Fuglerud, S.S.; Milenko, K.; Ellingsen, R.; Aksnes, A.; Hjelme, D.R. Feasibility of supercontinuum sources for use in glucose sensing by absorption spectroscopy. In Proceedings of the European Conference on Biomedical Optics 2019, Munich, Germany, 23–25 June 2019; p. 11073_13. [Google Scholar]
  133. Quarto, G.; Farina, A.; Pifferi, A.; Taroni, P.; Miniati, M. Time-resolved optical spectroscopy of the chest: Is it possible to probe the lung? In Proceedings of the European Conference on Biomedical Optics 2013, Munich, Germany, 12–16 May 2013; Taroni, P., Dehghani, H., Eds.; SPIE: Washington, DC, USA, 2013; p. 87990Q. [Google Scholar]
  134. Vignal, C.; Boumans, T.; Montcel, B.; Ramstein, S.; Verhoye, M.; Van Audekerke, J.; Mathevon, N.; Van der Linden, A.; Mottin, S. Measuring brain hemodynamic changes in a songbird: Responses to hypercapnia measured with functional MRI and near-infrared spectroscopy. Phys. Med. Biol. 2008, 53, 2457–2470. [Google Scholar] [CrossRef] [PubMed]
  135. Wabnitz, H.; Jelzow, A.; Mazurenka, M.; Steinkellner, O.; Macdonald, R.; Pifferi, A.; Torricelli, A.; Contini, D.; Zucchelli, L.M.G.; Spinelli, L.; et al. Performance assessment of time-domain optical brain imagers: A multi-laboratory study. In Proceedings of the SPIE BiOS, San Francisco, CA, USA, 2–7 February 2013; Volume 8583, pp. 85830L–85830L–14. [Google Scholar]
  136. Ferocino, E.; Di Sciacca, G.; Di Sieno, L.; Dalla Mora, A.; Pifferi, A.; Arridge, S.; Martelli, F.; Taroni, P.; Farina, A. Spectral approach to time domain diffuse optical tomography for breast cancer: Validation on meat phantoms. In Proceedings of the European Conference on Biomedical Optics 2019, Munich, Germany, 23–25 June 2019; p. 11074_7. [Google Scholar]
  137. Diop, M.; St. Lawrence, K. Improving the depth sensitivity of time-resolved measurements by extracting the distribution of times-of-flight. Biomed. Opt. Express 2013, 4, 447. [Google Scholar] [CrossRef] [PubMed]
  138. Nachabé, R.; Evers, D.J.; Hendriks, B.H.W.; Lucassen, G.W.; van der Voort, M.; Rutgers, E.J.; Peeters, M.-J.V.; Van der Hage, J.A.; Oldenburg, H.S.; Wesseling, J.; et al. Diagnosis of breast cancer using diffuse optical spectroscopy from 500 to 1600 nm: Comparison of classification methods. J. Biomed. Opt. 2011, 16, 087010. [Google Scholar] [CrossRef]
  139. Pifferi, A.; Farina, A.; Torricelli, A.; Quarto, G.; Cubeddu, R.; Taronia, P. Review: Time-domain broadband near infrared spectroscopy of the female breast: A focused review from basic principles to future perspectives. J. Near Infrared Spectrosc. 2012, 20, 223. [Google Scholar] [CrossRef] [Green Version]
  140. Lindner, C.; Mora, M.; Farzam, P.; Squarcia, M.; Johansson, J.; Weigel, U.M.; Halperin, I.; Hanzu, F.A.; Durduran, T. Diffuse optical characterization of the healthy human thyroid tissue and two pathological case studies. PLoS ONE 2016, 11, e0147851. [Google Scholar] [CrossRef]
  141. Lange, F.; Tachtsidis, I. Clinical brain monitoring with time domain NIRS: A review and future perspectives. Appl. Sci. 2019, 9, 1612. [Google Scholar] [CrossRef] [Green Version]
  142. Buttafava, M.; Martinenghi, E.; Tamborini, D.; Contini, D.; Mora, A.D.; Renna, M.; Torricelli, A.; Pifferi, A.; Zappa, F.; Tosi, A. A Compact Two-Wavelength Time-Domain NIRS System Based on SiPM and Pulsed Diode Lasers. IEEE Photonics J. 2017, 9, 1–14. [Google Scholar] [CrossRef] [Green Version]
  143. Bérubé-Lauzière, Y.; Crotti, M.; Boucher, S.; Ettehadi, S.; Pichette, J.; Rech, I. Prospects on Time-Domain Diffuse Optical Tomography Based on Time-Correlated Single Photon Counting for Small Animal Imaging. J. Spectrosc. 2016, 2016, 1–23. [Google Scholar] [CrossRef] [Green Version]
  144. Kaynezhad, P.; Mitra, S.; Bale, G.; Bauer, C.; Lingam, I.; Meehan, C.; Avdic-Belltheus, A.; Martinello, K.A.; Bainbridge, A.; Robertson, N.J.; et al. Quantification of the severity of hypoxic-ischemic brain injury in a neonatal preclinical model using measurements of cytochrome-c-oxidase from a miniature broadband-near-infrared spectroscopy system. Neurophotonics 2019, 6, 1. [Google Scholar] [CrossRef]
  145. Lange, F.; Bale, G.; Kaynezhad, P.; Pollock, R.D.; Stevenson, A.; Tachtsidis, I. Broadband NIRS Cerebral Evaluation of the Hemodynamic and Oxidative State of Cytochrome-c-Oxidase Responses to +Gz Acceleration in Healthy Volunteers. In Oxygen Transport to Tissue XLI; Springer: Cham, Switzerland, 2020; Volume 1232. [Google Scholar]
  146. Wu, J.; Zheng, G.; Liu, X.; Qiu, J. Near-infrared laser driven white light continuum generation: Materials, photophysical behaviours and applications. Chem. Soc. Rev. 2020, 49, 3461–3483. [Google Scholar] [CrossRef]
  147. Alayed, M.; Deen, M. Time-Resolved Diffuse Optical Spectroscopy and Imaging Using Solid-State Detectors: Characteristics, Present Status, and Research Challenges. Sensors 2017, 17, 2115. [Google Scholar] [CrossRef] [Green Version]
  148. Dalla Mora, A.; Di Sieno, L.; Re, R.; Pifferi, A.; Contini, D. Time-Gated Single-Photon Detection in Time-Domain Diffuse Optics: A Review. Appl. Sci. 2020, 10, 1101. [Google Scholar] [CrossRef] [Green Version]
  149. Mora, A.D.; Contini, D.; Arridge, S.; Martelli, F.; Tosi, A.; Boso, G.; Farina, A.; Durduran, T.; Martinenghi, E.; Torricelli, A.; et al. Towards next-generation time-domain diffuse optics for extreme depth penetration and sensitivity. Biomed. Opt. Express 2015, 6, 1749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Renna, M.; Buttafava, M.; Behera, A.; Zanoletti, M.; Di Sieno, L.; Mora, A.D.; Contini, D.; Tosi, A. Eight-Wavelength, Dual Detection Channel Instrument for Near-Infrared Time-Resolved Diffuse Optical Spectroscopy. IEEE J. Sel. Top. Quantum Electron. 2019, 25, 1–11. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Example of commercial SC laser sources and their typical spectra. (a) Fianium source, (b) Leukos source, (c) FYLA source, (d) YSL source. (e) Schematic of the classical triangular cladding single-core photonic crystal fiber in which light is guided in a solid core embedded in a triangular lattice of air holes. The fiber structure is determined by the hole-size, d, and the hole-pitch, Λ. Like standard fibers, the PCF is coated with a high index polymer for protection and to strip off cladding-modes. Extracted from the application note on supercontinuum generation in photonic crystal fibers by NKT. (f) Example of full spectrum generated by a SC laser, showing the width of the broadband emission and the central peak wavelength that is a residual of the single-mode laser used for SC generation. The spectrum is taken from the SuperK COMPACT SC laser manufactured by NKT Photonics (https://www.nktphotonics.com/lasers-fibers/product/superk-compact-supercontinuum-lasers/, accessed date: 1 November 2021).
Figure 1. Example of commercial SC laser sources and their typical spectra. (a) Fianium source, (b) Leukos source, (c) FYLA source, (d) YSL source. (e) Schematic of the classical triangular cladding single-core photonic crystal fiber in which light is guided in a solid core embedded in a triangular lattice of air holes. The fiber structure is determined by the hole-size, d, and the hole-pitch, Λ. Like standard fibers, the PCF is coated with a high index polymer for protection and to strip off cladding-modes. Extracted from the application note on supercontinuum generation in photonic crystal fibers by NKT. (f) Example of full spectrum generated by a SC laser, showing the width of the broadband emission and the central peak wavelength that is a residual of the single-mode laser used for SC generation. The spectrum is taken from the SuperK COMPACT SC laser manufactured by NKT Photonics (https://www.nktphotonics.com/lasers-fibers/product/superk-compact-supercontinuum-lasers/, accessed date: 1 November 2021).
Applsci 11 04616 g001
Figure 2. Pictures and schematic of various instruments based on SCL as a source. (a) TD-NIRS instrument aimed at broadband characterization of tissues. Figure reproduced from [23]. (b) TD-NIRS instrument aimed at monitoring both oxygenation and metabolism functional changes. Figure reproduced from [24]. (c) Preclinical instrument aimed at monitoring the exposed brain cortex of rodents. (reproduced with permission from [25], Copyright IEEE, 2021). (d) DOT system based on a SPAD camera aimed at functional brain monitoring in neonates. Figure reproduced from [26,27].
Figure 2. Pictures and schematic of various instruments based on SCL as a source. (a) TD-NIRS instrument aimed at broadband characterization of tissues. Figure reproduced from [23]. (b) TD-NIRS instrument aimed at monitoring both oxygenation and metabolism functional changes. Figure reproduced from [24]. (c) Preclinical instrument aimed at monitoring the exposed brain cortex of rodents. (reproduced with permission from [25], Copyright IEEE, 2021). (d) DOT system based on a SPAD camera aimed at functional brain monitoring in neonates. Figure reproduced from [26,27].
Applsci 11 04616 g002
Figure 3. Example of typical data extracted form system based on SCL as sources. (a) absorption and reduce scattering coefficient of several tissues acquired the TD-NIRS system developed in ref [23]. Figure reproduced from [23]. (b) 2D maps of the changes in optical properties of a phantom using MONSTIR II. (reproduced with permission from [28], Copyright AIP Publishing, 2021). (c) Changes in [HbO2] (red), [HHb] (blue) and [oxCCO](green), in the occipital lobe, acquired with the MAESTROS system [24], during a typical brain activation. Data presented at the fNIRS conference in Tokyo in 2018. (d) Raw image and reconstructed variation in [HbO2], [HHb], and [oxCCO] acquired on the exposed cortex during anoxia. (reproduced with permission from [25], Copyright IEEE, 2021).
Figure 3. Example of typical data extracted form system based on SCL as sources. (a) absorption and reduce scattering coefficient of several tissues acquired the TD-NIRS system developed in ref [23]. Figure reproduced from [23]. (b) 2D maps of the changes in optical properties of a phantom using MONSTIR II. (reproduced with permission from [28], Copyright AIP Publishing, 2021). (c) Changes in [HbO2] (red), [HHb] (blue) and [oxCCO](green), in the occipital lobe, acquired with the MAESTROS system [24], during a typical brain activation. Data presented at the fNIRS conference in Tokyo in 2018. (d) Raw image and reconstructed variation in [HbO2], [HHb], and [oxCCO] acquired on the exposed cortex during anoxia. (reproduced with permission from [25], Copyright IEEE, 2021).
Applsci 11 04616 g003
Figure 4. Summary of the literature review results. (a) Graph showing the number of publications using SCL in biomedical diffuse optics since the first study in 1993. (b) Summary of the main focus of the papers reviewed.
Figure 4. Summary of the literature review results. (a) Graph showing the number of publications using SCL in biomedical diffuse optics since the first study in 1993. (b) Summary of the main focus of the papers reviewed.
Applsci 11 04616 g004
Table 1. List of the SC laser-based systems used in the publications reviewed, with their principal characteristics. SC: SuperContinuum, FWHM: Full Widht at Half Maximum, CCD: Charge-Coupled Device, ICCD: Intensified Charge-Coupled Device, PMT: PhotoMultiplier Tube, SPAD: Single-Photon Avalanche Diode, SiPM: Silicon PhotoMultiplier, HPM: Hybrid Photodetector Module, sCMOS: scientific Complementary Metal–Oxide–Semiconductor, NIR: Near InfraRed.
Table 1. List of the SC laser-based systems used in the publications reviewed, with their principal characteristics. SC: SuperContinuum, FWHM: Full Widht at Half Maximum, CCD: Charge-Coupled Device, ICCD: Intensified Charge-Coupled Device, PMT: PhotoMultiplier Tube, SPAD: Single-Photon Avalanche Diode, SiPM: Silicon PhotoMultiplier, HPM: Hybrid Photodetector Module, sCMOS: scientific Complementary Metal–Oxide–Semiconductor, NIR: Near InfraRed.
System IDGroupYearSC Laser Type/ModelSC Laser System CharacteristicsSystem Detection Characteristics
Spectral CapacityPower (mW)Repetition Rate (kHz)Pulse Width (ps)Detector TypeN° of Channels
1Lund Institute of Technology, Lund, Sweden1993In-houseWhite light centred at 792 nm10000.01200CCD cameraNA
2Politecnico di Milano, Milan, Italy2004In-house550–1100 nm4085,00050–100PMT16
3Lund Institute of Technology, Lund, Sweden2004In-house500–1200 nmNA80,000100Streak cameraNA
4Université Jean Monnet, Saint-Etienne, France.2005In-house450–950 nm11170Streak cameraNA
5Politecnico di Milano, Milan, Italy2006In-house550–1050 nm100NA60–140PMT32
6Politecnico di Milano, Milan, Italy2007SC450, Fianium120 bands (600–1000 nm)
FWHM = 5–20 nm
2600 (total)20,00010SPAD1
7Lund University, Lund, Sweden2009SC500, Fianium650–1400 nmNA80,00050PMT2
8Northeastern University, Boston, United States2010Koheras SuperK, NKT Photonics550–850 nmNA80,00030PMT16
9National Optics Institute, Québec, Canada2010SC400, FianiumFiltered at 660 nmNA40,00090PMT1
10Physikalisch-Technische Bundesanstalt, Berlin, Germany2011Fianium (model NA)Tuned at 690 nmNA20,000100SPAD1
11University of California, Irvine, United States2012SC450, Fianium680–850 nm200020,0003PMT2
12Institute of Biocybernetics and Biomedical Engineering Polish Academy of Sciences, Warsaw, Poland2012SC450-4 (Fianium)685–860, 16 bandsNA40,000NAPMT16
13Politecnico di Milano, Milan, Italy2012SC450, Fianium1100–1700 nm
FWHM = 6.6–20.7 nm
6000 (total)40,00010SPAD1
14Massachusetts General Hospital, Boston, United States2013SC600-8, Fianium680, 710, 747, 760, 800, 820, 830, 840 nm9000 (total)60,000NAICCD175
15Politecnico di Milano, Milan, Italy2010SuperK Extreme, NKT Photonics600–1100 m5000 (total)2000–80,00010PMT1
16Katholieke Universiteit Leuven, Leuven, Belgium2012SC450-4, Fianium450–2400 nm4000 (total)NANAPINdiode1
17University College London, London, United Kingdom2014SC450, Fianium690, 750, 800, 850 nm4.540,0004PMT32
18Politecnico di Milano, Milan, Italy2014SC500-6, FianiumTuned at 800 nmNA40,500NAHPM1
19CREATIS, Université de Lyon, Lyon, France2015WhiteLase Micro, Fianium500–1000 nm102000NAICCD8
20Politecnico di Milano, Milan, Italy2015SC450, Fianium620 nm (40-nm bandwidth)NA80,000NASPAD1
21Politecnico di Milano, Milan, Italy2015SC450, Fianium750 nm (5-nm bandwidth)NA40,000NASPAD1
22Politecnico di Milano, Milan, Italy2015SC450, Fianium690 nmNA40,000NASiPM1
23The City College of New York, New York, United States2015STM-2000-IR, Leukos400–2500 nm0.5/nmNANAIR-CCD1
24Politecnico di Milano, Milan, Italy2015SC450, Fianium600–1350 nmNA40,000NASiPM and
PMT
2
25Politecnico di Milano, Milan, Italy2016SC500-6, Fianium760, 860 nm3040,500NASPAD1
26Physikalisch-Technische Bundesanstalt, Berlin, Germany2016SC500-6, FianiumTuned at 650 nmNA40,500100SPAD1
27Politecnico di Milano, Milan, Italy2017Fianium (model NA)750, 800, 850 nmNA40,000NASPAD2
28State Key Laboratory of Precision Measuring Technology and Instruments, Tianjin University, Tianjin, China2018SC46 (YSL photonics)1100–1350 nm (12 bands)8000 (Total)10–80,000NAInGaAs photodiode1
29Rensselaer Polytechnique Institute,
Troy, United States
2018MaiTai, Spectra Physics (1) SC Pro, YSL Photonics (2)(1) 690–1040 nm
FWHM = 15 nm
(2) 400–2200 nm
(1) NA
(2) 4000 (total)
(1) 80,000
(2) 25,000
(1) 100
(2) 150–200
PMT and
NIR camera
2
30Institute of Biocybernetics and Biomedical Engineering Polish Academy of Sciences, Warsaw, Poland2018SC450-4, Fianium650–850 nm, 16 bands of 12.5 nmNA40,000–80,000NAPMT1
31University of Pennsylvania, Philadelphia, United States2018SuperK Extreme, NKT Photonics730, 750, 786, 810, 830, 850 nmNA78,0005PMT2
32Physikalisch-Technische Bundesanstalt, Berlin, Germany2019FIU-15 PP, NKT PhotonicsNANANANASPAD and HPM2
33University of Science and Technology, Trondheim, Norway2019SuperK Compact, NKT Photonics (1) SCT 500, Fyla (2)NANA20 (1)
20,000 (2)
NAInGaAs detector2
34University of California, Irvine, United States2019SC400-2, Fianium1000 bands (580–950 nm) FWHM = 17.25 nm (mean)2000 (total)NANAsCMOS cameraNA
35University College London, London, United Kingdom2019SC480-6, Fianium16 bands (650–1100 nm) FWHM = 2–3 nm6000 (total)60,0004PMT4
36Physikalisch-Technische Bundesanstalt, Berlin, Germany2019SC500-6, FianiumTuned at 800 nmNA40,500NAHPM1
37Institute of Biocybernetics and Biomedical Engineering Polish Academy of Sciences, Warsaw, Poland2020FIU-15 PP, NKT PhotonicsTuned at 760 nmNA39,000NAHPM1
38University College London, London, United Kingdom2021WhiteLase Micro, Fianium600, 630, 665, 784, 800, 818, 835, 851, 868, 881, 894 nm (FWHM = 6–11 nm)0.06–0.11NANAsCMOS cameraNA
39Biomedical Optics Research Laboratory, University Hospital Zurich and University of Zurich, Switzerland2020SuperK ExtremeEXR-15 (NKT)Tuned at 800 nm6000 (total)80,000NASPAD
camera
NA
Table 3. Main relative advantages and drawbacks of single wavelength laser diode sources and SCL as function as the number of wavelengths used for TD systems. (+ advantages /− disadvantages). FW: Few Wavelength, HNW: High number of Wavelengths.
Table 3. Main relative advantages and drawbacks of single wavelength laser diode sources and SCL as function as the number of wavelengths used for TD systems. (+ advantages /− disadvantages). FW: Few Wavelength, HNW: High number of Wavelengths.
Single Wavelength SourcesSupercontinuum Laser
Few WavelengthsHigh number of WavelengthsFew WavelengthsHigh Number of Wavelengths
Cost++−−−−+
Stability++
Compactness++−−−−+
Wavelength ChoiceLimitedLimitedUnlimitedUnlimited
Overall+++−−−−++
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lange, F.; Giannoni, L.; Tachtsidis, I. The Use of Supercontinuum Laser Sources in Biomedical Diffuse Optics: Unlocking the Power of Multispectral Imaging. Appl. Sci. 2021, 11, 4616. https://doi.org/10.3390/app11104616

AMA Style

Lange F, Giannoni L, Tachtsidis I. The Use of Supercontinuum Laser Sources in Biomedical Diffuse Optics: Unlocking the Power of Multispectral Imaging. Applied Sciences. 2021; 11(10):4616. https://doi.org/10.3390/app11104616

Chicago/Turabian Style

Lange, Frédéric, Luca Giannoni, and Ilias Tachtsidis. 2021. "The Use of Supercontinuum Laser Sources in Biomedical Diffuse Optics: Unlocking the Power of Multispectral Imaging" Applied Sciences 11, no. 10: 4616. https://doi.org/10.3390/app11104616

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop