Open Access
3 August 2022 Noninvasive hemoglobin sensing and imaging: optical tools for disease diagnosis
Author Affiliations +
Abstract

Significance: Measurement and imaging of hemoglobin oxygenation are used extensively in the detection and diagnosis of disease; however, the applied instruments vary widely in their depth of imaging, spatiotemporal resolution, sensitivity, accuracy, complexity, physical size, and cost. The wide variation in available instrumentation can make it challenging for end users to select the appropriate tools for their application and to understand the relative limitations of different methods.

Aim: We aim to provide a systematic overview of the field of hemoglobin imaging and sensing.

Approach: We reviewed the sensing and imaging methods used to analyze hemoglobin oxygenation, including pulse oximetry, spectral reflectance imaging, diffuse optical imaging, spectroscopic optical coherence tomography, photoacoustic imaging, and diffuse correlation spectroscopy.

Results: We compared and contrasted the ability of different methods to determine hemoglobin biomarkers such as oxygenation while considering factors that influence their practical application.

Conclusions: We highlight key limitations in the current state-of-the-art and make suggestions for routes to advance the clinical use and interpretation of hemoglobin oxygenation information.

1.

Introduction

Optical-imaging biomarkers are defined characteristics measured with an optical imaging modality to indicate normal biological or pathological processes. Optical-imaging biomarkers can help researchers better understand disease development and give clinicians the ability to diagnose and treat diseases in patients.1,2

Based on the absorption of light by hemoglobin, optical imaging biomarkers such as hemoglobin concentration, oxygen saturation, and blood flow can be measured with a range of instruments for clinical disease evaluation. Hemoglobin is a protein in blood that transports oxygen to organs. Oxygen plays a vital role in cellular aerobic respiration, where it reacts with glucose to form adenosine triphosphate, water (H2O), and carbon dioxide (CO2), essential for maintaining healthy tissue and blood vessels. Hemoglobin oxygenation is often used as a vital sign; low oxygenation at the level of the organism can indicate a systemic disease, such as chronic obstructive pulmonary disease and apnea.3,4 Poor oxygenation in a particular organ or tissue can be symptomatic of an insult due to injury or illness, such as diabetes, skin trauma, rheumatic disease, or cancer.1,58

Noninvasive, low-cost, safe, and portable methods based on optics for extracting hemoglobin-derived biomarkers have become vital tools in patient management that can be applied in real time at the bedside. Despite widespread use of methods such as pulse oximetry, the uptake of newer technologies that go beyond point measurements remains relatively limited; however, there are many promising tools in development, ranging from 3D volumetric imaging of vascular architecture to spatially-resolved functional images of tissue oxygenation. Being less expensive and more portable in general than conventional radiological imaging methods, these have the potential to impact patient care in a wide range of debilitating illnesses, ranging from rheumatoid and vascular diseases to neurodegenerative diseases and cancer.

Here, we review these noninvasive methods for quantifying hemoglobin-derived biomarkers, including pulse oximetry, as commonly used in clinical practice worldwide, together with promising tools emerging in the research setting for imaging. The relative strengths and weaknesses of different methods are considered according to the application, grouped by mode of operation, including single-point detection, superficial imaging (up to 1- to 2-mm depth); and deep tissue imaging. We compare techniques based on the technology used, analysis methods, and current research or clinical applications; we then highlight limitations that would benefit from future research.

2.

Impact of Tissue Properties on Optical Measurement of Hemoglobin Biomarkers

2.1.

Biology of Human Blood

Human blood consists of plasma (about 55 vol.%) and cells (45 vol.%) in which 99% of the cells are red blood cells (RBCs) and the remaining 1% are leukocytes and thrombocytes.9,10 Plasma is a complex composition of dissolved ions (electrolytes), lipids, sugars, and proteins.10 RBCs, also known as erythrocytes, have a flat biconcave shape and a mean volume of 90  μm,3 9,11 and they contain about 30 pg of hemoglobin, a globular metalloprotein responsible for oxygen transport throughout the body.9,11 Hemoglobin concentrations range from 134 to 173  g/L in whole blood and 299 to 357  g/L in RBCs, varying according to age, gender, and health status. For example, anemia, cancer, or hereditary hemochromatosis decrease hemoglobin levels in the blood.9,12

Each hemoglobin molecule contains four heme groups that can be bound to oxygen; the unbound state is referred to as deoxygenated hemoglobin, Hb, and the saturated bound state is considered oxygenated and denoted by HbO2.10 Sometimes the oxygen saturation of arterial blood, SaO2, is differentiated from that of peripheral blood, StO2, because arterial blood oxygenation should be the same throughout the body. In contrast, peripheral blood has varying oxygenation levels as oxygen is absorbed from the blood by peripheral systems.11,13,14 Hemoglobin is also able to bind to other molecules forming carboxyhemoglobin, which arises during carbon monoxide inhalation;15,16 sulfhemoglobin, which arises due to the irreversible binding of sulphur in the presence of sulfonamides;17 and carbaminohemoglobin, which results from the binding of hemoglobin in venous blood to carbon dioxide.18 A further variant state of hemoglobin is methemoglobin,1921 in which hemoglobin binds iron in the Fe3+ state (unlike normal hemoglobin that binds Fe2+), which prevents the binding of oxygen. Methemoglobin occurs naturally in blood at 1% to 2% concentration,15 but it can be elevated due to side effects of medication or environmental factors.22 Finally, genetic variants of hemoglobin, such as hemoglobin S, which causes sickle cell anemia, can influence hemoglobin structure and binding properties.23 Myoglobin, a chromophore commonly found in muscles, binds oxygen with a higher affinity than hemoglobin.24 Myoglobin has a similar spectral response to hemoglobin and may be found in the bloodstream following muscle injury.24,25

2.2.

Optical Properties of Biological Tissue

Tissue is a complex turbid medium composed of different cell types and protein-rich extracellular matrix, which strongly impact the propagation of light.11,26,27 Absorption is the transformation of light energy to some other form of energy, such as heat, sound or fluorescence, as light traverses tissue and is quantified by the absorption coefficient, μa (cm1). Absorption is the primary optical interaction that is exploited to measure hemoglobin biomarkers. In turbid media, scattering is a major contributor to light attenuation and can confound attempts to measure hemoglobin absorption because blood is a highly scattering liquid with strong anisotropy. Scattering refers to a change in the direction of light propagation and is quantified by the scattering coefficient, μs (cm1), together with directional factors such as the scattering phase function and anisotropy factor, which further relate to the tissue refractive index.

2.2.1.

Optical properties of RBCs, hemoglobin, and its derivatives

The absorption coefficient of hemoglobin is a function of wavelength and the binding state (Fig. 1).32 Hemoglobin is usually oxygen-bound, whereas other variants mentioned above can modify the absorption spectrum and should be considered if relevant to the pathology being assessed because they can confound the measurement.10,13,23,33,34 It should be noted that the spectra of hemoglobin and its variety of physiologically relevant bound states are often measured after extracting the hemoglobin protein from RBCs, so they do not consider variation that arises due to scattering from different RBC geometries and orientations.

Fig. 1

The optical absorption of hemoglobin and associated variants. Representative spectra are shown for oxygenated,28 deoxygenated,28 carbamino,18 carboxy (visible29 and NIR19,20), methemoglobin (visible30 and NIR19,20), and sulfhemoglobin.31

JBO_27_8_080901_f001.png

In the absence of shear stress, human RBCs are biconcave discs with a diameter of 7 to 8  μm, maximal thickness of 2 to 3  μm, and minimal thickness of 0.8 to 1.5  μm.35,36 Concentrations of Hb within an RBC are high, on the order of 300 to 360  mg/mL, and the total refractive index of the cell is well approximated by that of pure hemoglobin through the application of the Kramers–Kronig relations on the total absorption spectrum.37,38 When light is incident upon a single RBC, scatter from the near and far membrane interfaces leads to a characteristic oscillatory scattering spectrum and phase function.36,39 The oscillatory spectral shape makes oximetry of a single RBC impossible without a priori knowledge of the precise shape and orientation of the RBC; however, averaging over many cells with random orientation can smooth this oscillation, which permits measurement of oxygen saturation of RBCs in capillaries.39,40 In addition, hematocrit, the volume fraction of RBCs in whole blood, the presence of blood plasma, and other factors affect the absorption spectra.18,4144

The most common hemoglobin-derived optical imaging biomarkers are total hemoglobin (referred to as THb hereafter) and oxygen saturation (referred to as sO2 hereafter). THb is often evaluated using a single wavelength absorption measurement taken at an isosbestic point of HbO2 and Hb (i.e., when their absorption coefficients are equal). sO2 requires an absorption measurement to be made at multiple wavelengths (at least 2), usually spanning regions where either HbO2 and Hb dominate the absorption properties. Data are then often analyzed by applying multivariate statistical approaches for spectral unmixing45 to extract the sO2 value. The absorption coefficients of HbO2 and Hb are related through sO2 to the overall optical absorption μa as11,45

Eq. (1)

sO2=[HbO2][HbO2]+[Hb],

Eq. (2)

μa(λ)=cHb{sO2100  μaHbO2(λ)+(1sO2100)μaHb(λ)},
where cHb is the concentration of hemoglobin in the tissue.

In addition to hemoglobin, many other molecules interact with light depending on their concentration and distribution throughout the tissue,46 which can disrupt light propagation and also introduce spectral coloring at depth in tissue, confounding attempts to measure THb and sO2. Furthermore, the scattering and refractive index properties of blood are affected by hemoglobin concentration, erythrocyte volume, shape, and aggregation, each of which can be modified in disease. Unlike the absorption coefficient, the anisotropy and the scattering coefficients of blood are not dependent on the changes in oxygenation.9,47

Hemoglobin-derived biomarkers often rely on variations in the absorption or scattering by hemoglobin molecules or RBCs, and the hardware used to make these measurements varies, including the use of LEDs or lasers for illumination with optical sensors (both arrays and point sensors) or ultrasound sensors for detection. The methods described in this review are summarized in Table 1.

Table 1

Overview of noninvasive hemoglobin monitoring and imaging.

TechnologyClinical application(s)Spatial resolutionBenefitsLimitationsClinical status and future potentialReferences
Pulse oximetryUsed to identify hypoxemia in a range of healthcare settings.Single point measurement with no spatial resolution.•Real-time monitoring of arterial saturation.
•Extremely simple and quick to use in a clinical setting.
•It can be used at the patient’s bedside.
•Single point location measurement.
•It can be inaccurate due to calibration assumptions
•Errors are associated with variations in hemoglobin and poor perfusion on the tissue measured.
Commonly used in primary through to tertiary care.
Potential for increased deployment at-home through wearables and low-cost devices.
13,4849.50.51.52.53.54.55.
56.57.58.59
Nailfold capillaroscopyDiagnosing and monitoring systemic scleroderma and other arthritic conditions.Spatial resolution varies from 0.1 to 10  μm laterally dependent upon the imaging NA, with corresponding imaging focal depth (Rayleigh range).•Real-time imaging of capillaries and blood flow is relatively easy.
•Typically, the design is portable, so it can be used at the bedside if needed.
•Lack of standardization around the quantification of capillary parameters.
•Measures only structural, not functional, information about the capillaries.
•Restricted to only imaging the nailfold capillaries.
Used in tertiary care. Emerging methods include quantification of blood flow, blood cell counts, and oxygenation imaging.5,6061.62.63.64.65.66.67.
68.69.70
Spectral reflectance imaging (MSI and HSI)Narrowband endoscopic imaging.Spatial resolution varies from 5 to 400  μm depending on the application because most systems use lenses for magnification. Reflectance signal restricted to superficial to 200  μm of tissue due to scattering.•A versatile method that can image multiple hemoglobin biomarkers as well as other proteins of interest.
•Relatively high spatial resolution is possible.
•Processing of images can be complicated and is not always possible in real time
•Snapshot methods are fast but sacrifice spectral and spatial resolution
•Scanning methods are slow but have a higher spectral resolution.
Commonly used in tertiary care for endoscopic surveillance.
Otherwise, used in research and small-scale clinical trials with promising applications in cancer detection, skin lesions, e.g., burns, and surgical tissue health.
2,6,8,30,7172.73.74.75.
76.77.78.79.80.81.82.83.
84.85
PAIBreast lesion evaluation.PAM can have a spatial resolution of 30  μm with an imaging depth of around 2 to 6 mm.
PAT can have a resolution of 200  μm with a penetration depth 2 to 3 cm.
•Can assess the relative hemoglobin saturation together with other biomarkers at depth while maintaining a reasonable spatial resolution.
•It is a versatile technique scaling spatial resolution with depth of imaging required.
•Trade-off between resolution and depth.
•Requires acoustic contact between the tissue and detectors.
•Systems can be costly, though LEDs can be used at the expense of imaging quality.
•The use of high-power pulsed light can present additional safety considerations to ensure that stray light does not damage patients or clinicians’ eyes
Large scale clinical trials (n>2000) for the diagnosis of breast cancer.
Small scale clinical trials or proof of concept for other applications, such as severity assessment in Crohn’s disease and dermatological conditions.
32,8687.88.89.90.91.92.93.
94.95.96.97.98.99.100.101.102.
103.104.105.106.107.108.109
Spectroscopic OCTStructural OCT is a clinical standard of care for retinal imaging.
Spectroscopic OCT not yet clinically approved.
Lateral resolution determined by illumination optics, typically tens of μm but as low as 1 to 2  μm is achievable.
Axial resolution determined by source/detector bandwidth, on the order of tens of μm down to 1  μm. Penetration depth is limited to 1  mm in medium-scattering tissue.
Real-time tomographic imaging with high resolution ideal for resolving tissue layers, structural characteristics in 3D.•Optical scattering in tissue limits imaging depth to 1 to 2 mm for most tissues.
•Laser scanning of sample introduces potential for motion artifacts in image, ophthalmic visible OCT implementation limited by safety threshold and patient aversion.
Structural OCT widely used in primary and tertiary centers for ophthalmology, dermatology, and dentistry, also in tertiary centers for cardiology.
Spectroscopic OCT for oximetry in preclinical development.
39, 110111.112.113.114.115.116.
117.118.119.120.121.122
DOI or NIRSAssessment of brain activity (fNIRS or DOT).
Monitoring of tissue oxygenation (cerebral oximetry/NIRS).
1 to 30 mm spatial resolution; high resolutions are only possible at shallow imaging depths (<2  cm). Low-resolution imaging is possible up to 10 cm into the tissue.Capable of determining blood oxygenation in tissue and other chromophores such as melanin, lipids, cytochrome-c-oxidase, and water.•Relatively low resolution; obtaining high resolution requires prior knowledge of tissue composition and a high density of optodes.
•DOI techniques may be combined with other imaging modalities such as MRI or ultrasound to assess tissue composition, resulting in increased system cost.
•Computationally-intensive and time-consuming, resulting in limited real-time imaging (spectroscopy does not have this problem).
Clinical approval and small-scale clinical use of tissue oximetry/NIRS for assessment of brain oxygenation.
fNIRS/DOT used commonly as a research tool to monitor brain activity.
Small scale clinical trials or proof of concept for other applications, such as breast cancer diagnosis, and joint inflammation.
123124.125.126.127.128.129.
130.131.132.133.134

3.

Point Sensing of Hemoglobin sO2 through Pulse Oximetry

Pulse oximetry makes a localized measurement of arterial hemoglobin sO2. To make this measurement, the absorption of tissue is evaluated at two or more different wavelengths, selected according to where the absorption coefficients of Hb and HbO2 differ sufficiently for their ratio to be evaluated as a biomarker that can be correlated directly to sO2 [Eq. (1)].

3.1.

Clinical Applications and Research Studies

Pulse oximetry has been extensively reviewed elsewhere.13,48,49,135,136 Pulse oximetry is deployed in many medical applications, from at-home first aid to clinical intensive care units and surgical theaters13, and it has found particular utility in assessing hypoxemia in COVID-19 patients137 as nonspecialists with minimal training can efficiently operate pulse oximeters.49,138 Pulse oximetry research in the clinic focuses on its use in treating and diagnosing diseases, such as optimizing oxygenation of ventilated patients, screening neonates for congenital heart diseases, and monitoring patients with sleep apneas.50,51,139,140 Despite widespread use, it is also well established that pulse oximetry can suffer racial bias, which results in less accurate oxygenation readings for patients with more skin melanin content, a trait associated with darker skin. The impact of such bias is severe as it has been shown to result in less adequate medical treatment of such patients, meaning that it is important for clinicians to be aware of this limitation and it is also an important area for future research and development.141,142 Although this bias has been well known for some time, it has been increasingly studied as a result of COVID-19 and the increased clinical use of pulse oximeters to treat respiratory conditions.

3.2.

Technology

Light absorption in pulse oximetry is typically measured using alternating illumination by LEDs at two different wavelengths.135,143,144 Because the wavelength of the illuminating light is altered with time, oxygenation measurements are susceptible to motion artifacts, which change the area of tissue being illuminated and coupling to the tissue, resulting in inaccuracies. Commonly used wavelength pairs are 660 and 940 nm or 665 and 894 nm,13,5255 which are applied in two different modes.

  • Transmission: Tissue such as the finger, toe, or earlobe is illuminated, and the light transmitted through the tissue is detected by a sensor to determine the amount of light attenuated by the tissue [Fig. 2(a)].

  • Reflection: Tissue such as the finger, foot, or forehead is illuminated, and the amount of light reflected by the tissue and underlying bone is detected by a sensor and used to determine the amount of light that the tissue has absorbed. Reflection pulse oximetry tends to have a higher signal when there is low perfusion.13,147 In reflection pulse oximetry, it can be challenging to isolate the light that has gone deeper into tissue from light that has been scattered or reflected at the surface of the tissue [Fig. 2(b)].

Fig. 2

Pulse oximetry. Schematic illustration of pulse oximetry in the two different operation modes: (a) transmission and (b) reflection. The detected light is cyclic due to the pulsatile nature of blood in the peripheral vascular system. Both transmission and reflection modes have alternating components (AC) and direct components (DC). In tissue, the transmission and reflection of light vary based on the changes in absorption due to blood volume and oxygenation. That is R+A+T1, when R, A, and T are the normalized reflection, absorption, and transmission intensities, respectively. For this reason, in reflection pulse oximetry, the peak intensity of light will be off by half a cycle from that of the transmission cycle. Examples of pulse oximeter devices include (c) transmission-based devices widely used in a clinical setting. Reproduced with permission from Ref. 145. (d) Low-power devices in development that adhere to the skin and use flexible OLED illumination. Reproduced with permission from Ref. 51. (e) Battery-free pulse oximeters in development that use near field communication for power. Reproduced with permission from Ref. 146.

JBO_27_8_080901_f002.png

Evolving from the traditional fingertip pulse oximeters [Fig. 2(c)], the current development of the technology mainly targets wearable devices and focuses on low power usage [Fig. 2(d)], optimization of signal detection, reduction of motion artifacts, flexible illumination and detection [Fig. 2(e)], low-cost devices, miniaturization, and calibration techniques.51,5659

3.3.

Analysis

The theory of oximetry analysis has been extensively reviewed by Mackenzie and Harvey,148 so it will be only briefly introduced; readers are referred to the prior review for a more detailed description. The total extinction coefficient for blood is denoted as ε and related to SaO2 as13

Eq. (3)

ε=εOSaO2+εD(1SaO2).

Further analysis of the extinction coefficient is needed to isolate the signal from arterial blood because venous blood also absorbs the light, along with other chromophores that appear in the light path, such as melanin (in the skin). It is possible to exploit the cyclic nature of the extinction coefficient due to the pumping of blood by finding the ratio of the variable component (AC) and constant component (DC) at two different wavelengths [Figs. 2(a) and 2(b)], where the difference in light absorption is rather large13,48,135

Eq. (4)

R=(AC/DC)1(AC/DC)2.

The ratio R, also known as the modulation ratio, is then related to SaO2 through a calibration procedure using best-fit analysis according to the equation

Eq. (5)

SaO2=a+bR,
where the variables a and b are calculated for each device during testing, based on a linear regression between the modulation ratio and the SaO2 value.135,149 Calibration was originally performed with human volunteers, changing SaO2 values by limiting the oxygen in the air that they breathed from 70% to 100% SaO2, which determined the R values.150 These are valid across devices with the same design, which means that individual devices did not have to be calibrated. Calibration techniques have evolved, so volunteers are no longer required. For example, several devices simulate the circulatory system and finger using pumps to mimic the pulsatile flow of arterial blood and venous blood.12,151155 The system is then sealed off, and the oxygenation of the blood can be controlled by varying the oxygen content of the system. Alternatives include electrical simulators that emit light from an LED corresponding to the light detected by the sensor to mimic the light transmitted in a typical finger. This technique requires prior knowledge of the pulse oximeter being calibrated. Once in operation, pulse oximeters are rarely recalibrated.150

3.4.

Limitations

Pulse oximetry measurements typically have an error of 3% to 4% depending on the device and calibration used, which is actually sufficient to impact patient care in some cases.150,156 In addition, standard pulse oximeters cannot detect the presence of methemoglobin, carboxyhemoglobin, or hemoglobin mutations, although their presence and concentrations outside of the expected range will affect the oxygenation readings.15,23 Some targeted pulse oximeters are able to detect methemoglobin and carboxyhemoglobin, but they are usually used in specific scenarios in which high levels of these derivatives are expected due to exposure.157 Hemoglobin F, present in fetuses and infants under 6 months, has a greater oxygen affinity, which allows the fetus to absorb oxygen from the mother’s bloodstream;10 in infants, its presence can increase the error in pulse oximetry by a further 3% in addition to the typical errors.13 Additionally, when there is poor perfusion to tissues, pulse oximetry can be limited, and if there is not a significant pulse detected, the technique will have increased inaccuracies.55 Finally, pulse oximetry has been found to suffer racial bias in two large cohorts, in which black patients had nearly three times the frequency of occult hypoxemia not detected by pulse oximetry as white patients.141,142 Skin pigmentation leads to an overestimation of arterial oxygen saturation in dark-skinned individuals, which could seriously impact medical decision making and long-term outcomes.158 These limitations merit increased attention in research and development given the potential for long-term and widespread use of pulse oximetry in COVID-19 patient management and the interest in deployment of the technology in the wearable setting.

4.

Reflectance Imaging of Hemoglobin

Optical imaging of hemoglobin biomarkers requires the operator to build a spatially resolved map of hemoglobin absorption at multiple wavelengths, again exploiting the differential absorption coefficients of Hb and HbO2.26,71,159 Several methods can be used to achieve this, including point-scanning spectroscopy, multispectral imaging, and hyperspectral imaging [Fig. 3(a)]. The result is a 3D dataset ((x,y,λ))26,72,159,163 that can be subjected to multivariate analysis methods to extract from the measured spectra the concentrations of their contributing chromophores (e.g., Hb and HbO2), referred to as “endmembers” for unmixing.26,72,164167 From these multivariate analyses, biomarkers that relate to THb and sO2 can then be extracted.

Fig. 3

Spectral reflectance imaging. (a) Overview of spectral reflectance imaging methods.159 Point-scanning spectroscopy can be used to build spectral information using a standard spectrometer. Alternatively, a spectral camera can be used to collect either a limited number of wavelengths (multispectral, typically <10 spectral bands) or a more continuous spectrum (hyperspectral). (b) Endoscopy images of the esophagus with (i) RGB imaging and (ii) narrowband imaging, which improves the contrast of the blood vessels. Reproduced with permission from Ref. 160. (c) Endoscopy of a porcine esophagus to determine tissue viability with 24 spectral bands from 460 to 690 nm (spectral resolution of 10 nm) using a slit hyperspectral imaging and fiber bundle probe and the resulting (i) reconstructed RGB image and (ii) unmixed oxygenation. Reproduced with permission from Ref. 161. (d) Hypoxia of tumors can be imaged using a liquid crystal tunable filter in conjunction with a CCD; this is demonstrated in mouse tumors; (i) and (iii) light microscopy of tumor vasculature in a dorsal skin window chamber, and the additional information of hemoglobin saturation is shown in (ii) and (iv) illustrating low oxygen saturation of the tumors. Reproduced with permission from Ref. 162.

JBO_27_8_080901_f003.png

4.1.

Clinical Applications and Research Studies

A widespread application of reflectance hemoglobin imaging is in gastrointestinal endoscopy. Virtual chromoendoscopy methods adapt the light source of the endoscopy to focus on two wavelength bands (415 and 540 nm), where hemoglobin absorbs strongly (Fig. 1), thus providing a high contrast morphological image of the tissue vasculature to the clinician.2,168 Narrowband imaging is the most widely established of these methods and meets the ASGE thresholds for targeting biopsies when imaging patients with Barrett’s esophagus for early signs of cancer.169 More recently, clinical research studies have demonstrated that, by expanding the number of wavelengths captured in endoscopy,72,73,170 it is possible to derive hemoglobin biomarkers of THb and sO2 from spectral information to classify disease status;171 however, further clinical study is needed to demonstrate efficacy.

Capillaroscopy is another reflectance-based imaging technique; it is used to image the blood vessels in the finger nailfold to diagnose disease, particularly to identify rheumatic diseases such as systemic scleroderma. Capillaries are microblood vessels from which oxygen and other nutrients are exchanged with the surrounding cells. In the finger nailfold, the capillaries are oriented in loops parallel to the skin, allowing for full visualization at high resolution in reflectance imaging mode. Capillary walls, formed from a single layer of endothelial cells lining the vessels, can rarely be detected during capillaroscopy, whereas the RBC column is visible, and morphological features associated with the capillary can be measured using monochrome, narrowband, and RGB imaging.60 Capillary blood flow in the finger ranges in velocity from 0.67 to 4  mm/s depending on physiological factors and the cyclic nature of perfusion.172174 Morphological dysfunction of the capillaries is easily identified using the current techniques, but current methods do not make oximetry measurements related to this dysfunction.

Reflectance-based oximetry imaging has been widely explored in retinal imaging because the retina is one of the most metabolically active tissues in the human body.175 Commercial retinal oximeters are applied to fundus cameras and use dual-wavelength illumination, akin to pulse oximetry, to acquire images simultaneously at one isosbestic wavelength and one sensitive to HbO2. Abnormal retinal oxygenation has been shown to detect diabetic retinopathy, age-related macular degeneration, and glaucoma.74,75,176 In addition, the retina has similar vascular properties to the brain, making it a perfect window for understanding and diagnosing neurological disease in addition to ocular disease.75 Nonetheless, retinal oximetry has yet to find routine clinical application, limited primarily by the impact of fundus pigmentation and vessel size on quantification.75 Although some hyperspectral imaging technologies have been explored in an attempt to overcome these limitations, to the best of our knowledge, they have not found clinical application.

In smaller scale studies, clinical trials of reflectance-based hemoglobin imaging have been applied in areas from the skin to the brain. For example, the hemodynamic response of the human cortex has been visualized during open-cranium surgery using hyperspectral imaging combined with multivariate analysis.76 Evaluation of hemoglobin oxygenation and melanin content in the face has been of interest for dermatological treatments such as the detection of skin cancer, assessment of scar healing, and evaluation of skin thickness.177180 Spectral imaging of hemoglobin in burns,8 wounds,6 and bruises30,77 has been explored to assess the progress of healing in a quantitative manner. Moreover, spectral imaging of tissue sO2 has found application in intraoperative imaging.161,181 Postoperative imaging can also provide clinicians with information on how tissue is healing such as following breast reconstruction in which water, hemoglobin concentration, and oxygen saturation are key indicators for tissue perfusion, an important factor in recovery.182

4.2.

Technology

The simplest imaging oximetry methods include two or three wavelengths akin to pulse oximetry, which can be applied using sequential illumination by the target wavelength bands and imaging with a single camera or simultaneous illumination of all wavelengths (e.g., with broadband illumination) and capture using a spectrally resolved method, such as image splitting through band pass filters applied in front of two cameras. Expanding the spectral range of wavelengths to capture more spectral features of Hb or HbO2 (in the visible but also near-infrared range183) requires more complex hardware.

There are three main types of multi or hyperspectral systems used in medical imaging that can be applied for unmixing oxygenation of tissue.78,79,159 In spatial scanning, a spectrograph records the spectral dimension (λ) while being scanned across a sample. A one-dimensional spectrograph may be point-scanned or a two-dimensional (2D) spectrograph may be line scanned. The approach sacrifices temporal resolution and hence requires minimal movement of the sample to avoid motion artifacts. In spectral scanning, a 2D camera records the spatial dimensions (x,y), while the spectral dimension is scanned, e.g., by changing the illumination wavelength or by filtering the imaging light path (e.g., with a filter wheel or tunable filter). This also has a scanning time associated with it, so it requires minimal movement to prevent spectral artifacts.

Finally, in snapshot methods, a system outputs spectral and spatial dimensions (x,y,λ) simultaneously without scanning.78,79 Snapshot systems may use beam splitting with dichroic filters, volume holographic optical element splitters, image replicating spectrometers, or multispectral filter arrays in the imaging path. Snapshot methods avoid motion artifacts, which makes them an exciting prospect for clinical application; however, they often exhibit poor optical efficiency. They also require a trade-off between spatial and spectral resolution, although this is less problematic for hemoglobin measurements, in which the target spectra are well characterized, and extensive evidence exists for the application of optical measurements. Nonetheless, optimization of the target wavelengths for snapshot imagers can substantially improve their performance.166,184187

In addition to these intensity-based imaging methods, spatial frequency domain imaging (SFDI) can be used to interrogate hemoglobin-related biomarkers by exploiting modulated illumination of the tissue. SFDI typically uses sinusoidal patterns and extracts the demodulation of these patterns reflected from tissue to calculate absorption and scattering properties; 123,188 low frequencies are more sensitive to absorption whereas high frequencies are more sensitive to scattering.123 Multiwavelength illumination189 can then be used to calculate hemoglobin content and oxygenation,190,191 e.g., in monitoring of peripheral circulation and vascular diseases,192 for which there are FDA-cleared devices, as well as ulcers,193 burns,194 or tumor margin detection.195,196 SFDI provides relatively high-resolution images, but conventional methods can be sensitive to motion artifacts188 and computational processing can limit the rate of image display. More recent reports have shown that it is possible to overcome these limitations using single snapshot of optical properties methods, which can achieve video-rate imaging.197

4.3.

Analysis

Two- or three-wavelength imaging methods may be viewed qualitatively and interpreted by the operator, as in narrowband imaging, or processed to output quantitative THb or sO2 biomarkers in a manner similar to pulse oximetry, through calculation of image ratios and calibration of the results. For spectral imaging methods, analysis can be time-intensive due to the large amount of data collected, which can be problematic for clinical translation in which the real-time display of biomarker data is often desired. Analysis methods vary depending on the biomarkers targeted and the type of tissue imaged from the simplest techniques such as linear spectral unmixing198,199 to more complex methods such as multivariate analysis and machine learning.26,72,164,165 Linear spectral unmixing determines the type and concentration of chromophores present based on input reference spectra for oxy and deoxy hemoglobin, from which oxygen saturation can be calculated.76,170,200,201 Spectral signatures can also be used directly in classification of disease status, e.g., cancer.171,181,202,203 Sometimes, a combination of classification and unmixing techniques can produce the optimal results, allowing for data corrections to be applied in certain tissue types.26,76,166,170 Similar methods are also used in depth-resolved imaging, but data may need to be corrected based on the imaging depth and the associated level of optical absorption and scattering. Machine learning methods have shown promise in this regard, enabling more accurate determination of hemoglobin oxygenation, particularly at depth, than classic linear spectral unmixing.204

4.4.

Limitations

Imaging tools for the assessment of hemoglobin can be subject to the same limitations as pulse oximetry. In addition, constraints that presently prevent clinical adoption of reflectance-based spectral imaging include cost, reliability of THb and sO2 measurements, clinical evidence for sensitivity and specificity of THb and sO2 in the diseases shown to be of interest in small-scale studies, and the need to process data in real time.2,71,167 These challenges are common in the clinical translation of optical imaging biomarkers,2 though fortunately, in the case of hemoglobin biomarkers, many devices have already navigated the pathway to the clinic, enabling initial studies to be undertaken.

5.

Depth-Resolved Imaging

A key limitation of pulse oximetry and reflectance-based imaging is their inability to provide depth-resolved imaging of hemoglobin biomarkers. Two modalities are available to determine THb and sO2 in tissue up to and beyond depths of 1 cm: photoacoustic imaging (PAI) and diffuse optical imaging (DOI); both have been evaluated in clinical trials and are at different stages of clinical adoption. PAI exploits the generation of acoustic waves by the absorption of pulsed light by hemoglobin to create deep tissue volumetric maps using pulsed illumination and ultrasound detectors [Fig. 4(a)].8688 DOI measures the properties of light scattering in tissue to generate absorption maps using synchronized illumination and photodiode-based detection [Fig. 4(b)].124126 At more restricted depths, below 1 cm, spectroscopic optical coherence tomography (OCT) is also applicable, combining broadband illumination with spectrally-resolved interferometric detection. Although these methods have been widely explored in the clinical research setting, they are only just beginning to find routine application in the clinic for patient management.

Fig. 4

Principles of depth-resolved imaging. (a) In photoacoustic imaging, the absorption of light pulses generates a broadband acoustic wave detected at the tissue surface by an ultrasound transducer. (b) Photoacoustic imaging of oxygenation of the finger in combination with ultrasound to image the veins and arteries. Reproduced with permission from Ref. 205. (c) In DOI (and DCS techniques), illuminated light is scattered in tissue collected by an offset optical detector at the tissue surface. (d) DOI data acquired from the human finger is processed to quantify oxygenation, hemoglobin concentration, and water. Reproduced with permission from Ref. 206. (e) In OCT, coherent light illuminates the tissue, and the light that reflects at interfaces is collected and combined with a reference arm, so interference occurs; from this interference, depth-resolved images of the absorption and scattering properties of tissue can be resolved. (f) Oxygen resolved spectroscopic OCT on mice brains illustrating how the fraction of inspired oxygen (FiO2) affects the oxygenation of the arteries and veins in the brain. Reproduced with permission from Ref. 207.

JBO_27_8_080901_f004.png

5.1.

Photoacoustic Imaging (PAI)

5.1.1.

Clinical research studies

By far, the most explored clinical PAI application is human breast cancer detection, extensively reviewed elsewhere,208,209 due to the enhanced angiogenesis of breast cancers compared with background breast parenchyma and the scalability of PAI geometry allowing for a broad view of the area [Fig. 5(a)].86,89,208,209,213,214 Multicenter clinical trials have recently been concluded covering >2000 women; these established the ability of PAI to increase the specificity of ultrasound imaging using a real-time map of relative Hb and HbO2.213,215 PAI has also been explored in other cancer types, considering that neoangiogenesis is a hallmark of cancer, including thyroid,216 prostate,215 and melanoma among, others.217 Beyond applications in cancer, PAI has found a wide range of potential applications in which depth-resolved information is required, e.g., in endoscopic procedures;90,91 for evaluation of inflammation, such as arthritic joints,88,92 foot ulcers,93 and Crohn’s disease;218 vascular imaging;94 and for the guidance of interventional procedures, such as in fetal placentas.95

Fig. 5

Tomographic imaging of the human breast for cancer detection (a) PAI of sO2 in breast with infiltrating ductal carcinoma (IDC); S-factor was defined to account for system accuracy and fluence compensation. Reproduced with permission from Ref. 210. (b) DOI of breast IDC (indicated by the red box) resolves sO2, THb, H2O, lipid, concentrations of which serve to highlight the tumor. Reproduced with permission from Ref. 211. (c) DCS of blood flow relative to an ultrasound image of low-grade carcinoma; the tumor is circled in yellow. These images are referenced to positions s1 and s2 to compare the ultrasound, 3D reconstruction, and cross-section. Reproduced with permission from Ref. 212.

JBO_27_8_080901_f005.png

5.1.2.

Technology

Pulsed illumination is required to excite PAI signals. Typically, tunable pulsed lasers with nanosecond pulse durations have been used; however, pulsed laser diodes and LEDs have emerged recently as viable alternatives.219 The generated acoustic waves are detected by ultrasound transducers, which may be single-element, linear, or curvilinear arrays or in a spherical arrangement, depending on the system’s geometry. The type of transducer and associated center frequency or bandwidth is usually governed by the application, depending on the absorber size, laser pulse width, and required imaging depth.220 PAI can be deployed in different geometries, including tomography, mesoscopy, and microscopy. Tomography systems have found the most widespread clinical application as they provide an adequate field of view and spatial resolution for imaging of hemoglobin in deep tissue such as the breast; mesoscopy systems have also been applied clinically to visualize vascular network architectures in the skin given their limited penetration depths.221

5.1.3.

Analysis

The acoustic wave generated in response to pulsed optical illumination depends on the absorption properties of tissue according to

Eq. (13)

p0(z)=ΓμaF0eμ0z,
where p0 is the initial pressure, Γ is the Gruneisen parameter, F0 is the initial fluence, μ0 is a constant, and z is the depth of the tissue.96 PAIs are reconstructed using a range of beamforming methodologies, akin to ultrasound imaging.94,96 3D tomographic images can be reconstructed by combining the temporal and spatial information collected, which is often achieved analytically using a simple back-projection inversion or numerically using model-based methods.96 Images reconstructed from data acquired at several wavelengths can then be subjected to the same multivariate analysis methods described in Sec. 4 for spectral unmixing. However, frame-to-frame coregistration may be needed to avoid spatial or spectral corruption due to motion.

5.1.4.

Limitations

The attainable depth of PAI depends on the optical and acoustic attenuation of the sample. In soft tissue, acoustic attenuation scales as a function of ultrasound frequency, so at low frequencies of a few MHz, optical attenuation tends to dominate and is the constraining factor for imaging depth.96 For spatial resolution, the constraining factor is the bandwidth of the acoustic wave, usually limited by the acoustic attenuation of soft tissue and the frequency response of the detecting transducer.97 The latter is particularly important for imaging more superficial features when the bandwidth of the signal can extend to 100 MHz and beyond, for which there is a limited availability of high-performance transducers. The maximum acoustic frequency transmitted decreases with depth, meaning that typically systems that operate at higher penetration depths have lower spatial resolutions than those designed for shallow imaging.97

A key challenge for PAI is biomarker quantification. During reconstruction, a number of assumptions are made; these include the speed of sound in tissue, transducer impulse response, detection bandwidth, and continuous sampling.96 If these assumptions break down, for example, due to heterogeneities in tissue due to air cavities, there will be distortions in the image. Furthermore, when evaluating biomarkers such as THb and sO2 the nature of light propagation in tissue can lead to distortions in the spectral properties of the illumination as a function of depth. Although some methods have been explored to compensate for such “spectral coloring,”94,222 they often break down in the complex scenarios found in human tissue and have yet to be validated in a clinical setting. Finally, as with all hemoglobin sensing and imaging methods, calibration of the extracted biomarkers is vital. PAI calibration and clinical quality assurance methods are still under development, particularly through a community-led effort.98

5.2.

Diffuse Optical Spectroscopy and Imaging

5.2.1.

Clinical research studies

Diffuse optical spectroscopy (DOS) is commonly referred to as near-infrared spectroscopy (NIRS) because it uses light in the near-infrared range; the term functional NIRS (fNIRS) is also commonly used but is usually restricted to applications monitoring functional responses to stimuli in the brain via neurovascular coupling. Quantifying and monitoring changes in oxygenation of blood in the brain has found many applications that range from understanding seizures223 to detecting brain damage.7 Unlike reflectance hemoglobin imaging [Figs. 6(a)6(d)], in which an open cranium is required (see Sec. 4), fNIRS typically achieves imaging depths of up to 15 mm through the skull [Figs. 6(e) and 6(f)], which covers the outer cerebral cortex in healthy adults.127,223225 fNIRS imaging of the brain to identify intracranial hematomas due to brain trauma has been clinically approved by the FDA recently; however, it has yet to be widely deployed in clinical settings.7

Fig. 6

Hemoglobin imaging of the human brain. (a)–(d) Reflectance spectral images of the brain in an adult undergoing epileptogenic tissue resection (a) reference RGB rendering, (b) change in oxygenated hemoglobin over a single timeframe, (c) change in deoxygenated hemoglobin over a single timeframe, and (d) change in total hemoglobin over a single timeframe. Reproduced with permission from Ref. 76. (e), (f) DOI of a neonate during a seizure: (e) changes in HbT concentration mapped throughout the onset of a seizure and (f) average changes in Hb, HbO2, and tHb postonset of the seizure. Reproduced with permission from Ref. 223.

JBO_27_8_080901_f006.png

Similar to PAI, DOI has also been deployed in clinical trials to detect cancer, particularly in the breast226,227 [Figs. 5(b) and 5(c)] and thyroid,127 where it has also been used to monitor response to therapy. DOI tends to be lower in spatial resolution than PAI (Fig. 5), but it can often resolve other biomarkers in addition to hemoglobin. PAI typically has a lower temporal resolution compared with DOI.228,229 In addition, DOI has been used to image muscle tissue such as the forearm, peripheral tissue, and joints.127,225 Several reviews have been published illustrating the importance of DOI.230233

5.2.2.

Technology

DOI techniques typically use near-infrared light in the 650 to 900 nm range.127,225,228,230 Below 650 nm, light can experience poor penetration due to the absorption of hemoglobin in superficial tissues, whereas above 920 nm, water absorption comes to dominate.225 Most DOI and fNIRS systems use two (or more) different wavelengths, one that is in the lower range below the NIR isosbestic point, such as 680, 695, 705, or 730 nm, and another that has a longer wavelength, such as 830 or 850 nm. The number of light sources and detectors varies depending on the system from 1 to >48, with some designed to allow for more light sources and detectors to enable customization of spatial resolution according to the target application.228,234

DOS is similar to DOI but provides single-point measurements of tissue properties such as hemoglobin and blood oxygen saturation similar to pulse oximetry.235 There are three main modalities of DOS/I: continuous wave (CW), time-domain (TD), and frequency-domain (FD) systems.127,225,228 CW systems detect changes in the intensity of illumination, making them relatively simple and inexpensive.236 Slight changes in surface coupling can affect the intensity measurements, resulting in poor reproducibility unless well controlled. TD systems correlate the time between emission and detection of photons to measure the photon flight time. Tissue scattering determines photon flight time, whereas absorption determines the overall intensity of photons reaching the detector. FD systems are based on similar principles to TD systems but instead measure the phase shift of the incoming light. Frequency domain systems are significantly less expensive than TD systems due to lower costs for the sensors and detectors. TD and FD systems can distinguish contaminating signals due to background illumination because these signals will be uncorrelated. Because TD and FD systems can separate out the effects of scattering and absorption, these techniques can quantify absolute concentrations of hemoglobin and its oxygenation, whereas CW systems typically measure the relative change in concentration but not the absolute quantities.

DOI can be deployed in different geometries, either topographic, with imaging of a single plane with limited depth information, or tomographic (DOT), including depth resolution, which allows for full 3D reconstruction of the pertinent properties of tissue. In transillumination techniques, the illumination and detection are on opposite sides of the tissue being imaged; however, this is limited to body parts with small radii. Measurements at multiple and overlapping source–detector separations can be used to create depth measurements and reconstruct a 3D image. In tomographic systems, the illumination and detection sensors are placed on the available surface to simultaneously measure the changes in illumination throughout the sample.

Finally, it is worth noting that diffuse correlation spectroscopy (DCS) is similar to DOI and DOS, but it uses an autocorrelation function through a combination of hardware and software to measure an index of blood flow in tissue [Fig. 5(c)].128 Due to the similarities in apparatus, sometimes DCS is combined with DOI or DOS systems128 to provide a spatially-resolved indication of blood flow as a complementary biomarker to THb or sO2.128 Blood flow in tissue can also be detected using laser speckle contrast imaging (LCSI), which looks at fluctuations in the speckle pattern reflected from tissue to determine the flow rate of the blood.237,238 Laser Doppler flowmetry (LDF) finds the flow rate and concentration of blood by quantifying the Doppler shift that causes spectral broadening of reflected light. 238,239 All three of these techniques can resolve flow rate, but they typically have relatively low spatial resolutions or are confined to single-point measurements. Recent advances, particularly in LCSI, are now reaching near real-time operation at higher spatial resolutions, benefitting from increased computing power available in portable systems.240

5.2.3.

Analysis

The analysis for DOI depends on the type of imaging system used. The measured signals can be converted into optical absorption maps by understanding the transport of light in tissue, which can be modeled using the radiative transport equation (RTE)127,225 or Monte Carlo methods. The RTE is an analytical approach that approximates Maxwell’s equations in diffuse media, assuming a constant refractive index. Solving the RTE is computationally expensive,225 but it can be sped up by exploiting symmetries or by making approximations. For example, the diffusion approximation assumes isotropic scattering,127 so it can be applied only in diffusive tissues; it does not work well in nondiffusive tissue, such as the cerebrospinal fluid that surrounds the brain or in anisotropic media, such as the skin and nervous system, where prior information is required to reach a solution. Numerical techniques, such as the finite element method, finite difference method, finite volume method, and boundary element method, can be used to solve the diffusion approximation. The use of prior information such as MRI, CT, or other imaging techniques can vastly improve DOI resolution by reducing the number of assumptions about the tissue structure that are made.225

Light transport can also be understood by forward modeling the propagation of light using Monte Carlo methods to calculate the propagation of photons through media. The accumulated forward modeling statistics can then be analyzed with respect to real data from DOI systems to address the inverse scattering problem for various anisotropic scattering media.225 Monte Carlo methods have traditionally also been computationally expensive, but they are now reaching higher speeds with deployment on GPU and cloud-based servers.241

5.2.4.

Limitations

DOI exploits the scattering properties of tissue, unlike other imaging techniques that are hampered by light scattering. Despite this, DOI still has limited spatial resolution (on the order of 1 mm230) and often requires anatomical priors from another modality such as MRI or CT for analysis, which constrains applications127,225,230,235 and adds cost and complexity to high-resolution DOI systems.225 Standardization of DOI systems for clinical translation is ongoing, e.g., through a project to characterize DOI systems using phantoms,242 particularly for breast cancer detection.243

5.3.

Spectroscopic OCT

5.3.1.

Clinical research studies

While OCT is primarily deployed for structural imaging, its spectrally-resolved detection can be harnessed for angiography and oximetry in vivo, albeit not yet in clinical practice. OCT is typically implemented in the NIR for structural assessment due to its greater penetration through tissue and availability of light sources with suitable coherence. Extraction of blood sO2 from OCT measurements in human retinal vasculature was first demonstrated in the NIR (800/850 nm),244 but it had limited precision due to weak absorption in this range. The development of visible OCT systems improved the available signal-to-noise ratio and hence precision, enabling measurements from single erythrocytes110,111 and high-resolution capillary oximetry in 3D.112,113,245 Advances in reconstruction algorithms and high-speed instrumentation have improved OCT angiography to a point at which it has found clinical use for high contrast imaging of retinal and dermal vasculature.246,247 Visible OCT measurement of retinal sO2 has been tested in humans114 alongside angiography,115 but it is still in development. Laser exposure limits and natural aversion restrict the usable power level for visible OCT in the eye compared with clinical NIR OCT, but continuing improvements in OCT technology have allowed for high-quality imaging and sO2 measurement.

5.3.2.

Technology

OCT is a noncontact imaging modality that can be considered an optical analog of ultrasound imaging.116 The distinguishing feature of OCT is the use of low-coherence interferometry to decouple the lateral and axial imaging resolution: lateral resolution is determined by the numerical aperture of the focusing optic, whereas axial resolution is determined by the temporal coherence of the laser used for imaging. Scanning OCT systems operate in the Fourier domain,248 where a broad spectrum of light interrogates the tissue and is then collected through spectrally-resolved detection for postprocessing and image reconstruction by an inverse Fourier transform. Within this class, there are two major varieties: spectral-domain OCT (SD-OCT), which uses broadband illumination and parallel detection with a spectrometer, and swept-source OCT (SS-OCT), which uses a high-speed spectral scanning source with single-channel detection for temporally-resolved spectral acquisition.

Recent advances in SS-OCT laser technology, including the tunable vertical-cavity source emitting laser, have enabled the realization of high-speed, robust, and compact instruments, which leads to greater imaging depths. Current SS-OCT systems operate exclusively in the NIR range due to swept laser availability; for imaging performance (and potential Hb oximetry) in the visible domain, SD-OCT systems using a visible spectrometer are required. In recent years, the high power and spatial coherence of newly available supercontinuum lasers have enabled good-quality visible OCT imaging in research systems,117,249,250 which could enable future developments toward clinical translation.

5.3.3.

Analysis

The raw data for SD-OCT are acquired from a high-speed spectrometer with line readout synchronized to a scanning mechanism, allowing for the mapping of each spectrum to a spatial position. The spectrum is normalized, filtered, and converted to a spatial reflectance profile, known as an A-line, through an inverse Fourier transform. For oximetry, the SD of this analysis must be narrowed, typically through a windowing function; although full spectral resolution of the OCT volume may be realized through the application of the short-time Fourier transform, the dependence of axial resolution on spectral bandwidth poses a necessary trade-off between spectral resolution and axial resolution in the spectroscopic OCT image. For sO2 measurement and spectral-contrast angiography, spectral windows are chosen to maximize the contrast of hemoglobin, typically in the range of 550 to 650 nm.112,251 The spectrally-resolved total extinction coefficient is measured through depth fitting of the spectroscopic OCT A-line signal to the Beer–Lambert law, from which the relative contributions of Hb and HbO2 can be unmixed to determine sO2.

5.3.4.

Limitations

OCT is highly versatile, being deployed for imaging the inner walls of blood vessels and luminal organs,252256 having an ultrawide field of view for scanning skin,257259 and with corrective lenses, compensating for ocular refraction in the ophthalmologic clinic. Nonetheless, there are several fundamental issues that are limiting its general adoption. First, the scanned acquisition adds instrumental complexity and can produce motion artifacts in patients. Although this is addressed in full-field OCT systems, these typically do not work with rough samples. Next, the spectral resolution of OCT imaging determines the maximum depth range that can be imaged, an aspect referred to as the sensitivity roll-off. Finally, because OCT is primarily sensitive to singly-scattered photons in tissue, the penetration depth of imaging is restricted to 1 to 2 mm in most human tissues. For this reason, large-scale clinical deployment of OCT has largely been limited to ophthalmology, dermatology, and cardiology, but creative advances in OCT probe and capsule technology will allow for continued in situ exploration of hemoglobin-related biomarkers from OCT in disease pathology throughout the body.256,260262

6.

Summary and Perspective

Moving beyond pulse oximetry to exploit the optical absorption of hemoglobin in imaging applications has shown significant promise in the clinic, with both superficial 2D and depth-resolved 3D implementations described in this review. Two wavelength THb and sO2 imaging are already widely used in endoscopic and ophthalmic applications, respectively, and have reached large-scale clinical trials in depth-resolved PAI. Conversely, reflectance-based spectral imaging remains largely exploratory.

6.1.

Trade Offs

Choosing the optimal hemoglobin imaging technique for a particular application involves consideration of several factors that often require trade-offs, including signal-to-noise ratio, spatial and temporal resolutions, target depth, and route to integration with existing clinical practice. Optical imaging techniques are ultimately restricted by the maximum permissible exposure at the illumination site, which places a fundamental limit on the signal-to-noise ratio available in the clinical setting. Some techniques are further restricted in the type of illumination used such as OCT, which requires coherent light, and PAI, which requires short, pulsed light; this can add to the complexity of safety considerations in the clinic.

Considering the factors of resolution and depth, reflectance-based imaging can achieve high spatiotemporal resolution in applications in which depth resolution is not vital, for example, with retinal, endoscopic, or intraoperative imaging. One could argue that depth-resolution will become increasingly prevalent with the emergence of more advanced solutions from spectroscopic OCT, PAI, and DOI, particularly as costs decrease. Nonetheless, depth-resolution typically implies a sacrifice of spatial or temporal resolution, which must be determined early in the discovery and development phase of the associated device. Furthermore, different approaches to achieving depth resolution have different strengths and weaknesses. PAI may suffer from shielding effects due to the absorption of overlying tissue and tends to resolve only larger vessels, whereas the use of multiple scattering by DOI enables better detection of capillary oxygenation, despite the overall poorer spatial resolution. This trade-off may explain why PAI has developed more rapidly in clinical translation for breast cancer detection, whereas DOI is more developed for imaging the brain. Given the common contrast source across many of the techniques described, it may also be desirable to combine multiple approaches in a single device for validation purposes or to provide views of the same tissue at different resolution scales, overcoming some of the limitations of the existing technologies.263 Combining two or more optical imaging techniques can be beneficial because they reduce the imaging limitations that arise as a result of a single technique. A good example of this is the enhanced perfusion oximetry system that combines diffuse reflectance spectroscopy with LDF to quantify blood oxygenation and flow rate for imaging of microvasculature and burns.264266

6.2.

Clinical Implementation

There are many factors to consider regarding the route to integration of new techniques into clinical practice. For example, if imaging is required to be non-contact, for example, in delicate targets like the eye, this may restrict the implementation of methods such as PAI or DOI, in which contact with the tissue is typically required. The pathway to clinical adoption may be smoother in the case of an existing optical imaging solution already being deployed. An obvious example is the large-scale deployment of OCT for ophthalmic applications, which provides a direct route for adoption of spectroscopic OCT in the community. Another factor in clinical translation is the assessment of the precision, accuracy, and bias of new biomarker measurements.2 These factors can be affected by both device operation and data interpretation, requiring standardization of data acquisition and careful consideration of any signal processing or analysis methods applied before presenting the data to the interpreter, whether this is a human or a machine. Because reflectance-based imaging techniques are usually less computationally intensive compared with the depth-resolved methods, the development of systems and software for real-time imaging is more easily attainable; this is seen most effectively in narrowband imaging of blood vessels in endoscopy.

Further clinical considerations arise later in the translational pathway when the question of biomarker efficacy in decision-making finally arises. Instrument prototypes are often used first in pilot clinical trials at a single clinical center to gather initial data for validation and may be subject to multiple design iterations at this stage. Having successfully passed the first translational gap, which may include CE marking or FDA approval for the device and multicenter clinical trials, the technique is then subject to advanced qualification and ongoing technical validation to determine clinical utility in the healthcare setting, whereby the measurement can be used in clinical decision making. These larger-scale clinical trials help to determine sources of variation that will influence the classification and diagnosis of disease, providing clinical evidence of the ability to change patient management. They also provide extensive reference data sets that can be used to improve interpretation, particularly when machine learning-based methods are involved.

Reflectance-based spectral imaging techniques are still largely in the earlier stages of development with first in-person trials, whereas fNIRS of the brain and PAI of the breast are being used in multiple centers as part of larger-scale clinical trials and DOI has found some level of adoption into the clinic.71,87,208,228,231,267 Data arising from these trials is extremely valuable and making annotated datasets open source for the community in the future will not only help accelerate the development of new algorithms but could also enhance our understanding of the biology of hemoglobin oxygenation in disease.

6.3.

Disease Monitoring

Hemoglobin imaging methods could find further applications in monitoring disease, to detect treatment efficacy or disease relapse. The noninvasive nature of these techniques can allow for continuous or periodic monitoring, for which there are several excellent examples that have been highlighted. Pulse oximetry can be applied to a patient for long periods so that clinicians can observe if there are any changes to overall arterial oxygenation. DOI and NIRS also can be used longitudinally to monitor changes in brain function to assess if there is improved brain activity.267 Furthermore, periodic monitoring of diseases such as scleroderma using nailfold capillaroscopy can indicate progression in the stage and severity of the disease, which may influence the treatment provided or indicate if further medical intervention is required.

Periodic monitoring can be applied in the short term, for monitoring wound or burn healing, or in the longer term in the context of endoscopic cancer surveillance in at-risk patient groups such as those with Barrett’s esophagus that are at increased risk of developing cancer. The frequency of monitoring applied is determined by the disease being observed and the rate at which change is expected, as well as by the training required for instrument use and data interpretation. For example, applying pulse oximetry is relatively quick with nonspecialists able to do it, and some more straightforward versions of capillaroscopy can be done using handheld devices. Conversely, endoscopies, OCT, DOI, and NIRS typically require training of specialist operators; hence it can be more expensive to conduct the procedures, meaning they are typically used for less frequent monitoring.

6.4.

Outlook

The acceptability and relevance of new hemoglobin sensing and imaging technologies to clinicians will be driven by various factors, including cost, complexity, and physical size of the systems, as well as the ease of use and data interpretation. The commonplace use of pulse oximetry means that the clinical community is already well aware of the use of systemic sO2 as a disease biomarker. Ongoing technological developments that lead to miniaturization of light sources, optical components, and cameras, as well as decreasing their cost, mitigate some of the technical limitations highlighted in this review. Advances in image processing, including convolutional neural networks, promise to aid in distilling rich datasets to actionable clinical information, enabling imaging systems to be more easily integrated into clinical care. Research questions remain regarding the sensitivity of hemoglobin sensing techniques in diverse populations and the diagnostic power of hemoglobin-derived biomarkers in the wide array of disease presentations included in this review, which will only be answered through comprehensive clinical trials. Hemoglobin imaging techniques add a new dimension of knowledge in a range of clinical settings, from capillaroscopy and endoscopy to intraoperative imaging; emerging technologies are well placed to further enhance these areas of existing clinical practice, but are also likely to contribute to the decentralization of healthcare to tertiary care centers and through the deployment of wearable technologies for self-monitoring in the home.

Disclosure

The authors have no conflicts of interest to declare.

Acknowledgments

M. T-W. acknowledges the financial support of the General Sir John Monash Foundation and the Cambridge Trust. GB was supported by the Gianna Angelopoulos Programme for Science Technology and Innovation. GS and SEB received funding from the EPSRC (EP/R003599/1) and CRUK (C9545/A29580).

References

1. 

J. P. B. O’Connor et al., “Imaging biomarker roadmap for cancer studies,” Nat. Rev. Clin. Oncol., 14 (3), 169 –186 (2017). https://doi.org/10.1038/nrclinonc.2016.162 Google Scholar

2. 

D. J. Waterhouse et al., “A roadmap for the clinical implementation of optical-imaging biomarkers,” Nat. Biomed. Eng., 3 (5), 339 –353 (2019). https://doi.org/10.1038/s41551-019-0392-5 Google Scholar

3. 

J. Levy et al., “Digital oximetry biomarkers for assessing respiratory function: standards of measurement, physiological interpretation, and clinical use,” NPJ Digit. Med., 4 (1), (2021). https://doi.org/10.1038/s41746-020-00373-5 Google Scholar

4. 

A. S. Scott, M. A. Baltzan and N. Wolkove, “Examination of pulse oximetry tracings to detect obstructive sleep apnea in patients with advanced chronic obstructive pulmonary disease,” Can. Respir. J., 21 (3), 171 –175 (2014). https://doi.org/10.1155/2014/948717 Google Scholar

5. 

G. Dinsdale et al., “State-of-the-art technologies provide new insights linking skin and blood vessel abnormalities in SSc-related disorders,” Microvasc. Res., 130 (April), 104006 (2020). https://doi.org/10.1016/j.mvr.2020.104006 MIVRA6 0026-2862 Google Scholar

6. 

A. Nouvong et al., “Evaluation of diabetic foot ulcer healing with hyperspectral imaging of oxyhemoglobin and deoxyhemoglobin,” Diabetes Care, 32 (11), 2056 –2061 (2009). https://doi.org/10.2337/dc08-2246 DICAD2 0149-5992 Google Scholar

7. 

C. S. Robertson et al., “Clinical evaluation of a portable near-infrared device for detection of traumatic intracranial hematomas,” J. Neurotrauma, 27 (9), 1597 –1604 (2010). https://doi.org/10.1089/neu.2010.1340 JNEUE4 0897-7151 Google Scholar

8. 

M. S. Chin et al., “Hyperspectral imaging for burn depth assessment in an animal model,” Plast. Reconstr. Surg. - Glob. Open, 3 (12), e591 –9 (2015). https://doi.org/10.1097/GOX.0000000000000558 Google Scholar

9. 

A. Roggan et al., “Optical properties of circulating human blood in the wavelength range 400-2500 nm,” J. Biomed. Opt., 4 (1), 36 –46 (1999). https://doi.org/10.1117/1.429919 JBOPFO 1083-3668 Google Scholar

10. 

M. L. Turgeon, Clinical Hematology: Theory and Procedures, 2nd ed.Little, Brown, Boston; London (1993). Google Scholar

11. 

I. J. Bigio and S. Fantini, “Overview of tissue optical properties,” Quantitative Biomedical Optics: Theory, Methods, and Applications, 19 –59 1st ed.Cambridge University Press(2016). Google Scholar

12. 

T. Vos et al., “Global, regional, and national incidence, prevalence, and years lived with disability for 310 diseases and injuries, 1990–2015: a systematic analysis for the Global Burden of Disease Study 2015,” Lancet, 388 (10053), 1545 –1602 (2016). https://doi.org/10.1016/S0140-6736(16)31678-6 LANCAO 0140-6736 Google Scholar

13. 

M. Nitzan, A. Romem and R. Koppel, “Pulse oximetry: fundamentals and technology update,” Med. Devices Evid. Res., 7 (1), 231 –239 (2014). https://doi.org/10.2147/MDER.S47319 Google Scholar

14. 

W. Shin, Y. D. Cha and G. Yoon, “ECG/PPG integer signal processing for a ubiquitous health monitoring system,” J. Med. Syst., 34 (5), 891 –898 (2010). https://doi.org/10.1007/s10916-009-9304-7 JMSYDA 0148-5598 Google Scholar

15. 

B. A. Darlow and C. J. Morley, “Oxygen saturation targeting and bronchopulmonary dysplasia,” Clin. Perinatol., 42 (4), 807 –823 (2015). https://doi.org/10.1016/j.clp.2015.08.008 CLPEDL 0095-5108 Google Scholar

16. 

N. W. Tietz and E. A. Fiereck, “The spectrophotometric measurement of carboxyhemoglobin,” Ann. Clin. Lab. Sci., 3 (1), 1388 –1393 (1973). ACLSCP 0091-7370 Google Scholar

17. 

L. Gharahbaghian, B. Massoudian and G. Dimassa, “Methemoglobinemia and sulfhemoglobinemia in two pediatric patients after ingestion of hydroxylamine sulfate,” West. J. Emerg. Med., 10 (3), 197 –201 (2009). Google Scholar

18. 

E. Dervieux, Q. Bodinier and W. Uhring, “Measuring hemoglobin spectra: searching for carbamino-hemoglobin,” J. Biomed. Opt., 25 (10), 105001 (2020). https://doi.org/10.1117/1.JBO.25.10.105001 Google Scholar

19. 

M. Van Gastel, S. Stuijk and G. De Haan, “Camera-based pulse-oximetry - validated risks and opportunities from theoretical analysis,” Biomed. Opt. Express, 9 (1), 102 (2018). https://doi.org/10.1364/BOE.9.000102 BOEICL 2156-7085 Google Scholar

20. 

J. K. Barton et al., “Cooperative phenomena in two-pulse, two-color laser photocoagulation of cutaneous blood vessels,” Photochem. Photobiol., 73 (6), 642 (2001). https://doi.org/10.1562/0031-8655(2001)0730642CPITPT2.0.CO2 PHCBAP 0031-8655 Google Scholar

21. 

L. L. Randeberg et al., “Methemoglobin formation during laser induced photothermolysis of vascular skin lesions,” Lasers Surg. Med., 34 (5), 414 –419 (2004). https://doi.org/10.1002/lsm.20042 LSMEDI 0196-8092 Google Scholar

22. 

S. R. David et al., “The blood blues: a review on methemoglobinemia,” J. Pharmacol. Pharmacother., 9 (1), 1 –5 (2018). https://doi.org/10.4103/jpp.JPP_79_17 Google Scholar

23. 

C. S. Thom et al., “Hemoglobin variants: biochemical properties and clinical correlates,” Cold Spring Harb. Perspect. Med., 3 (3), a011858 (2013). https://doi.org/10.1101/cshperspect.a011858 Google Scholar

24. 

L. Tofani et al., “Spectroscopic and interfacial properties of myoglobin/surfactant complexes,” Biophys. J., 87 (2), 1186 –1195 (2004). https://doi.org/10.1529/biophysj.104.041731 BIOJAU 0006-3495 Google Scholar

25. 

L. S. L. Arakaki, D. H. Burns and M. J. Kushmerick, “Accurate myoglobin oxygen saturation by optical spectroscopy measured in blood-perfused rat muscle,” Appl. Spectrosc., 61 (9), 978 –985 (2007). https://doi.org/10.1366/000370207781745928 APSPA4 0003-7028 Google Scholar

26. 

G. Lu and B. Fei, “Medical hyperspectral imaging: a review,” J. Biomed. Opt., 19 (1), 010901 (2014). https://doi.org/10.1117/1.JBO.19.1.010901 JBOPFO 1083-3668 Google Scholar

27. 

V. Tuchin, Tissue Optics: Light Scattering Methods and Instruments for Medical Diagnosis, 3rd ed.SPIE Press, Bellingham, Washington, USA (2015). Google Scholar

28. 

S. A. Prahl, “Tabulated molar extinction coefficient for hemoglobin in water,” (1998) http://omlc.ogi.edu/spectra/hemoglobin/summary.html Google Scholar

29. 

O. Siggaard-Andersen, B. Nørgaard-Pedersen and J. Rem, “Hemoglobin pigments spectrophotometric determination of oxy-, carboxy-, met-, and sulfhemoglobin in capillary blood,” Clin. Chim. Acta, 42 (1), 85 –100 (1972). https://doi.org/10.1016/0009-8981(72)90380-4 CCATAR 0009-8981 Google Scholar

30. 

L. L. Randeberg and J. Hernandez-Palacios, “Hyperspectral imaging of bruises in the SWIR spectral region,” Proc. SPIE, 8207 82070N (2012). https://doi.org/10.1117/12.909137 PSISDG 0277-786X Google Scholar

31. 

I. H. Yarynovska and A. I. Bilyi, “Absorption spectra of sulfhemoglobin derivates of human blood,” Proc. SPIE, 6094 60940G (2006). https://doi.org/10.1117/12.639597 PSISDG 0277-786X Google Scholar

32. 

J. Yao and L. V. Wang, “Sensitivity of photoacoustic microscopy,” Photoacoustics, 2 (2), 87 –101 (2014). https://doi.org/10.1016/j.pacs.2014.04.002 Google Scholar

33. 

W. G. Zijlstra, A. Buursma and W. P. Meeuwsen-van der Roest, “Absorption spectra of human fetal and adult oxyhemoglobin, de-oxyhemoglobin, carboxyhemoglobin, and methemoglobin,” Clin. Chem., 37 (9), 1633 –1638 (1991). https://doi.org/10.1093/clinchem/37.9.1633 Google Scholar

34. 

A. P. Harris et al., “Absorption characteristics of human fetal hemoglobin at wavelengths used in pulse oximetry,” J. Clin. Monit., 4 (3), 175 –177 (1988). https://doi.org/10.1007/BF01621812 JCMOEH 0748-1977 Google Scholar

35. 

K. A. Shapovalov, “The geometry and optical models of the erythrocyte,” Syst. Rev. Pharm., 11 (12), 1456 –1463 (2020). https://doi.org/10.31838/srp.2020.12.216 Google Scholar

36. 

M. Kinnunen et al., “Effect of the size and shape of a red blood cell on elastic light scattering properties at the single-cell level,” Biomed. Opt. Express, 2 (7), 1803 (2011). https://doi.org/10.1364/BOE.2.001803 BOEICL 2156-7085 Google Scholar

37. 

J. Gienger et al., “Refractive index of human red blood cells between 290 nm and 1100 nm determined by optical extinction measurements,” Sci. Rep., 9 4623 (2019). https://doi.org/10.1038/s41598-019-38767-5 SRCEC3 2045-2322 Google Scholar

38. 

D. J. Faber, F. J. Van Der Meer and M. C. G. Aalders, “Quantitative measurement of attenuation coefficients of weakly scattering media using optical coherence tomography,” Opt. Express, 12 (19), 4353 –592 (2004). https://doi.org/10.1364/OPEX.12.004353 OPEXFF 1094-4087 Google Scholar

39. 

R. Liu et al., “Theoretical model for optical oximetry at the capillary level: exploring hemoglobin oxygen saturation through backscattering of single red blood cells,” J. Biomed. Opt., 22 (2), 025002 (2017). https://doi.org/10.1117/1.JBO.22.2.025002 JBOPFO 1083-3668 Google Scholar

40. 

R. Liu et al., “Single capillary oximetry and tissue ultrastructural sensing by dual-band dual-scan inverse spectroscopic optical coherence tomography,” Light Sci. Appl., 7 (1), 57 (2018). https://doi.org/10.1038/s41377-018-0057-2 Google Scholar

41. 

N. Bosschaart et al., “A literature review and novel theoretical approach on the optical properties of whole blood,” Lasers Med. Sci., 29 (2), 453 –479 (2014). https://doi.org/10.1007/s10103-013-1446-7 Google Scholar

42. 

M. Meinke et al., “Optical properties of platelets and blood plasma and their influence on the optical behavior of whole blood in the visible to near infrared wavelength range,” J. Biomed. Opt., 12 (1), 014024 (2007). https://doi.org/10.1117/1.2435177 JBOPFO 1083-3668 Google Scholar

43. 

M. Meinke et al., “Empirical model functions to calculate hematocrit-dependent optical properties of human blood,” Appl. Opt., 46 (10), 1742 –1753 (2007). https://doi.org/10.1364/AO.46.001742 APOPAI 0003-6935 Google Scholar

44. 

M. Friebel et al., “Influence of oxygen saturation on the optical scattering properties of human red blood cells in the spectral range 250 to 2000 nm,” J. Biomed. Opt., 14 (3), 034001 (2009). https://doi.org/10.1117/1.3127200 JBOPFO 1083-3668 Google Scholar

45. 

T. Saito and H. Yamaguchi, “Optical imaging of hemoglobin oxygen saturation using a small number of spectral images for endoscopic imaging,” J. Biomed. Opt., 20 (12), 126011 (2015). https://doi.org/10.1117/1.JBO.20.12.126011 Google Scholar

46. 

S. L. Jacques, “Optical properties of biological tissues: a review,” Phys. Med. Biol., 58 (11), R37 –R61 (2013). https://doi.org/10.1088/0031-9155/58/11/R37 PHMBA7 0031-9155 Google Scholar

47. 

A. J. Welch, M. J. C. van Gemert and J. T. Walsh, “Basic Interactions of Light with Tissue,” Optical-Thermal Response of Laser-Irradiated Tissue, 13 –26 Springer Netherlands, Dordrecht (2011). Google Scholar

48. 

A. Reisner et al., “Utility of the photoplethysmogram in circulatory monitoring,” Anesthesiology, 108 (5), 950 –958 (2008). https://doi.org/10.1097/ALN.0b013e31816c89e1 ANESAV 0003-3022 Google Scholar

49. 

J. E. Sinex, “Pulse oximetry: principles and limitations,” Am. J. Emerg. Med., 17 (1), 59 –66 (1999). https://doi.org/10.1016/S0735-6757(99)90019-0 Google Scholar

50. 

G. Tusman, S. H. Bohm and F. Suarez-Sipmann, “Advanced uses of pulse oximetry for monitoring mechanically ventilated patients,” Anesth. Analg., 124 (1), 62 –71 (2017). https://doi.org/10.1213/ANE.0000000000001283 Google Scholar

51. 

H. Lee et al., “Toward all-day wearable health monitoring: an ultralow-power, reflective organic pulse oximetry sensing patch,” Sci. Adv., 4 (11), 1 –9 (2018). https://doi.org/10.1126/sciadv.aas9530 STAMCV 1468-6996 Google Scholar

52. 

S. Lopez, Pulse Oximeter Fundamentals and Design, 1 –39 Free Semiconductor Inc.(2012). Google Scholar

53. 

O. Y. Hay et al., “Pulse oximetry with two infrared wavelengths without calibration in extracted arterial blood,” Sensors (Switzerland), 18 (10), 3457 (2018). https://doi.org/10.3390/s18103457 Google Scholar

54. 

M. P. McEwen, G. P. Bull and K. J. Reynolds, “Vessel calibre and haemoglobin effects on pulse oximetry,” Physiol. Meas., 30 (9), 869 –883 (2009). https://doi.org/10.1088/0967-3334/30/9/001 PMEAE3 0967-3334 Google Scholar

55. 

E. Hill and M. D. Stoneham, “Practical applications of pulse oximetry,” Updat. Anaesth., 5 (11), 11 –15 (2000). Google Scholar

56. 

T. Y. Abay and P. A. Kyriacou, “Comparison of NIRS, laser Doppler flowmetry, photoplethysmography, and pulse oximetry during vascular occlusion challenges,” Physiol. Meas., 37 (4), 503 –514 (2016). https://doi.org/10.1088/0967-3334/37/4/503 PMEAE3 0967-3334 Google Scholar

57. 

C. Liu et al., “Optical fibre-based pulse oximetry sensor with contact force detection,” Sensors (Switzerland), 18 (11), 3632 (2018). https://doi.org/10.3390/s18113632 Google Scholar

58. 

W. Verkruysse et al., “Calibration of contactless pulse oximetry,” Anesth. Analg., 124 (1), 136 –145 (2017). https://doi.org/10.1213/ANE.0000000000001381 Google Scholar

59. 

J. Kim et al., “Miniaturized battery-free wireless systems for wearable pulse oximetry,” Adv. Funct. Mater., 27 (1), 1 –8 (2017). https://doi.org/10.1002/adfm.201604373 AFMDC6 1616-301X Google Scholar

60. 

S. N. Lambova and U. Müller-Ladner, “Nailfold capillaroscopy in systemic sclerosis – state of the art: the evolving knowledge about capillaroscopic abnormalities in systemic sclerosis,” J. Scleroderma Relat. Disord., 4 (3), 200 –211 (2019). https://doi.org/10.1177/2397198319833486 Google Scholar

61. 

M. V. Volkov et al., “Video capillaroscopy clarifies mechanism of the photoplethysmographic waveform appearance,” Sci. Rep., 7 13298 (2017). https://doi.org/10.1038/s41598-017-13552-4 SRCEC3 2045-2322 Google Scholar

62. 

J. Aguirre et al., “Assessing nailfold microvascular structure with ultra-wideband raster-scan optoacoustic mesoscopy,” Photoacoustics, 10 31 –37 (2018). https://doi.org/10.1016/j.pacs.2018.02.002 Google Scholar

63. 

A. K. Murray et al., “Noninvasive imaging techniques in the assessment of scleroderma spectrum disorders,” Arthritis Care Res., 61 (8), 1103 –1111 (2009). https://doi.org/10.1002/art.24645 Google Scholar

64. 

A. Karbalaie et al., “Practical issues in assessing nailfold capillaroscopic images: a summary,” Clin. Rheumatol., 38 (9), 2343 –2354 (2019). https://doi.org/10.1007/s10067-019-04644-9 Google Scholar

65. 

D. Paxton and J. D. Pauling, “Does nailfold capillaroscopy help predict future outcomes in systemic sclerosis? A systematic literature review,” Semin. Arthritis Rheum., 48 (3), 482 –494 (2018). https://doi.org/10.1016/j.semarthrit.2018.02.005 Google Scholar

66. 

M. Berks et al., “An automated system for detecting and measuring nailfold capillaries,” Lect. Notes Comput. Sci., 8673 658 –665 (2014). https://doi.org/10.1007/978-3-319-10404-1_82 LNCSD9 0302-9743 Google Scholar

67. 

M. E. Anderson et al., “Computerized nailfold video capillaroscopy - a new tool for assessment of Raynaud’s phenomenon,” J. Rheumatol., 32 (5), 841 –848 (2005). JORHE3 Google Scholar

68. 

G. N. McKay, N. Mohan and N. J. Durr, “Imaging human blood cells in vivo with oblique back-illumination capillaroscopy,” Biomed. Opt. Express, 11 (5), 2373 (2020). https://doi.org/10.1364/BOE.389088 BOEICL 2156-7085 Google Scholar

69. 

M. Michalska-Jakubus et al., “Plasma endothelial microparticles reflect the extent of capillaroscopic alterations and correlate with the severity of skin involvement in systemic sclerosis,” Microvasc. Res., 110 24 –31 (2017). https://doi.org/10.1016/j.mvr.2016.11.006 MIVRA6 0026-2862 Google Scholar

70. 

B. Ruaro et al., “Correlations between skin blood perfusion values and nailfold capillaroscopy scores in systemic sclerosis patients,” Microvasc. Res., 105 119 –124 (2016). https://doi.org/10.1016/j.mvr.2016.02.007 MIVRA6 0026-2862 Google Scholar

71. 

N. T. Clancy et al., “Surgical spectral imaging,” Med. Image Anal., 63 101699 (2020). https://doi.org/10.1016/j.media.2020.101699 Google Scholar

72. 

J. Yoon et al., “A clinically translatable hyperspectral endoscopy (HySE) system for imaging the gastrointestinal tract,” Nat. Commun., 10 1902 (2019). https://doi.org/10.1038/s41467-019-09484-4 NCAOBW 2041-1723 Google Scholar

73. 

A. S. Luthman et al., “Bimodal reflectance and fluorescence multispectral endoscopy based on spectrally resolving detector arrays,” J. Biomed. Opt., 24 (3), 031009 (2018). https://doi.org/10.1117/1.JBO.24.3.031009 JBOPFO 1083-3668 Google Scholar

74. 

R. Podlipec et al., “Characterization of blood coagulation dynamics and oxygenation in ex-vivo retinal vessels by fluorescence hyperspectral imaging,” J. Biophotonics, 13 (8), 1 –12 (2020). https://doi.org/10.1002/jbio.202000021 Google Scholar

75. 

X. Hadoux et al., “Non-invasive in vivo hyperspectral imaging of the retina for potential biomarker use in Alzheimer’s disease,” Nat. Commun., 10 49 (2019). https://doi.org/10.1038/s41467-019-12242-1 NCAOBW 2041-1723 Google Scholar

76. 

J. Pichette et al., “Intraoperative video-rate hemodynamic response assessment in human cortex using snapshot hyperspectral optical imaging,” Neurophotonics, 3 (4), 045003 (2016). https://doi.org/10.1117/1.NPh.3.4.045003 Google Scholar

77. 

L. C. Cancio, “Application of novel hyperspectral imaging technologies in combat casualty care,” Proc. SPIE, 7596 759605 (2010). https://doi.org/10.1117/12.846331 PSISDG 0277-786X Google Scholar

78. 

A. S. Luthman, Spectrally Resolved Detector Arrays for Multiplexed Biomedical Fluorescence Imaging, Springer International Publishing, Cham (2018). Google Scholar

79. 

S. Grusche, “Basic slit spectroscope reveals three-dimensional scenes through diagonal slices of hyperspectral cubes,” Appl. Opt., 53 (20), 4594 (2014). https://doi.org/10.1364/AO.53.004594 APOPAI 0003-6935 Google Scholar

80. 

M. Ewerlöf, M. Larsson and E. G. Salerud, “Spatial and temporal skin blood volume and saturation estimation using a multispectral snapshot imaging camera,” Proc. SPIE, 10068 1006814 (2017). https://doi.org/10.1117/12.2251928 PSISDG 0277-786X Google Scholar

81. 

N. Hagen and M. W. Kudenov, “Review of snapshot spectral imaging technologies,” Opt. Eng., 52 (9), 090901 (2013). https://doi.org/10.1117/1.OE.52.9.090901 Google Scholar

82. 

M. Torabzadeh et al., “Hyperspectral imaging in the spatial frequency domain with a supercontinuum source,” J. Biomed. Opt., 24 (7), 071614 (2019). https://doi.org/10.1117/1.JBO.24.7.071614 JBOPFO 1083-3668 Google Scholar

83. 

M. A. Wares et al., “Noninvasive evaluation of hemodynamics and light scattering property during two-stage mouse cutaneous carcinogenesis based on multispectral diffuse reflectance images at isosbestic wavelengths of hemoglobin,” J. Biomed. Opt., 24 (3), 031020 (2019). https://doi.org/10.1117/1.JBO.24.3.031020 JBOPFO 1083-3668 Google Scholar

84. 

D. J. Waterhouse et al., “Emerging optical methods for endoscopic surveillance of Barrett’s oesophagus,” Lancet Gastroenterol. Hepatol., 3 (5), 349 –362 (2018). https://doi.org/10.1016/S2468-1253(18)30030-X Google Scholar

85. 

T. Chiba et al., “Advanced multispectral image-processing endoscopy system for visualizing two-dimensional hemoglobin saturation and relative hemoglobin concentration,” Endosc. Int. Open, 7 (11), E1442 –E1447 (2019). https://doi.org/10.1055/a-0990-9189 Google Scholar

86. 

L. Lin et al., “Single-breath-hold photoacoustic computed tomography of the breast,” Nat. Commun., 9 2352 (2018). https://doi.org/10.1038/s41467-018-04576-z NCAOBW 2041-1723 Google Scholar

87. 

M. Li, Y. Tang and J. Yao, “Photoacoustic tomography of blood oxygenation: a mini review,” Photoacoustics, 10 65 –73 (2018). https://doi.org/10.1016/j.pacs.2018.05.001 Google Scholar

88. 

J. Jo et al., “A functional study of human inflammatory arthritis using photoacoustic imaging,” Sci. Rep., 7 15026 (2017). https://doi.org/10.1038/s41598-017-15147-5 SRCEC3 2045-2322 Google Scholar

89. 

S. Manohar and M. Dantuma, “Current and future trends in photoacoustic breast imaging,” Photoacoustics, 16 100134 (2019). https://doi.org/10.1016/j.pacs.2019.04.004 Google Scholar

90. 

P. K. Upputuri and M. Pramanik, “Recent advances toward preclinical and clinical translation of photoacoustic tomography: a review,” J. Biomed. Opt., 22 (4), 041006 (2016). https://doi.org/10.1117/1.JBO.22.4.041006 JBOPFO 1083-3668 Google Scholar

91. 

W. C. Chapman and M. Mutch, “Co-registered photoacoustic and ultrasound imaging of human colorectal cancer,” J. Biomed. Opt., 24 (12), 121913 (2019). https://doi.org/10.1117/1.jbo.24.12.121913 JBOPFO 1083-3668 Google Scholar

92. 

G. S. Sangha and C. J. Goergen, “Label-free photoacoustic and ultrasound imaging for murine atherosclerosis characterization,” APL Bioeng., 4 (2), 026102 (2020). https://doi.org/10.1063/1.5142728 Google Scholar

93. 

Y. Wang et al., “A portable three-dimensional photoacoustic tomography system for imaging of chronic foot ulcers,” Quantum Imaging Med. Surg., 9 (5), 799 –799 (2019). https://doi.org/10.21037/qims.2019.05.02 Google Scholar

94. 

V. Ntziachristos and D. Razansky, “Molecular imaging by means of multispectral optoacoustic tomography (MSOT),” Chem. Rev., 110 (5), 2783 –2794 (2010). https://doi.org/10.1021/cr9002566 CHREAY 0009-2665 Google Scholar

95. 

E. Maneas et al., “Photoacoustic imaging of the human placental vasculature,” J. Biophotonics, 13 (4), 1 –10 (2020). https://doi.org/10.1002/jbio.201900167 Google Scholar

96. 

M. Xu and L. V. Wang, “Photoacoustic imaging in biomedicine,” Rev. Sci. Instrum., 77 (4), 041101 (2006). https://doi.org/10.1063/1.2195024 RSINAK 0034-6748 Google Scholar

97. 

P. Beard, “Biomedical photoacoustic imaging,” Interface Focus, 1 (4), 602 –631 (2011). https://doi.org/10.1098/rsfs.2011.0028 Google Scholar

98. 

S. Bohndiek, “Addressing photoacoustics standards,” Nat. Photonics, 13 (5), 298 (2019). https://doi.org/10.1038/s41566-019-0417-3 NPAHBY 1749-4885 Google Scholar

99. 

G. Ku et al., “Multiple-bandwidth photoacoustic tomography,” Phys. Med. Biol., 49 (7), 1329 –1338 (2004). https://doi.org/10.1088/0031-9155/49/7/018 PHMBA7 0031-9155 Google Scholar

100. 

T. P. Nguyen et al., “Improved depth-of-field photoacoustic microscopy with a multifocal point transducer for biomedical imaging,” Sensors (Switzerland), 20 (7), 1 –18 (2020). https://doi.org/10.3390/s20072020 Google Scholar

101. 

C. Chu et al., “Multimodal photoacoustic imaging-guided regression of corneal neovascularization: a non-invasive and safe strategy,” Adv. Sci., 2000346 1 –7 (2020). https://doi.org/10.1002/advs.202000346 Google Scholar

102. 

M. Martinho Costa et al., “Quantitative photoacoustic imaging study of tumours in vivo: baseline variations in quantitative measurements,” Photoacoustics, 13 53 –65 (2019). https://doi.org/10.1016/j.pacs.2018.12.002 Google Scholar

103. 

P. Zhang et al., “High-resolution deep functional imaging of the whole mouse brain by photoacoustic computed tomography in vivo,” J. Biophotonics, 11 (1), 1 –6 (2018). https://doi.org/10.1002/jbio.201700024 Google Scholar

104. 

R. Haindl et al., “Functional optical coherence tomography and photoacoustic microscopy imaging for zebrafish larvae,” Biomed. Opt. Express, 11 (4), 2137 (2020). https://doi.org/10.1364/BOE.390410 BOEICL 2156-7085 Google Scholar

105. 

H. Leng et al., “Characterization of a fiber bundle-based real-time ultrasound/photoacoustic imaging system and its in vivo functional imaging applications,” Micromachines, 10 (12), 820 (2019). https://doi.org/10.3390/mi10120820 Google Scholar

106. 

A. Hariri et al., “The characterization of an economic and portable LED-based photoacoustic imaging system to facilitate molecular imaging,” Photoacoustics, 9 10 –20 (2018). https://doi.org/10.1016/j.pacs.2017.11.001 Google Scholar

107. 

Y. Junjie et al., “High-speed label-free functional photoacoustic microscopy of mouse brain in action,” Nat. Methods, 12 (4), 407 –410 (2015). https://doi.org/10.1038/nmeth.3336 1548-7091 Google Scholar

108. 

H. F. Zhang et al., “Functional photoacoustic microscopy for high-resolution and noninvasive in vivo imaging,” Nat. Biotechnol., 24 (7), 848 –851 (2006). https://doi.org/10.1038/nbt1220 NABIF9 1087-0156 Google Scholar

109. 

A. Taruttis and V. Ntziachristos, “Advances in real-time multispectral optoacoustic imaging and its applications,” Nat. Photonics, 9 (4), 219 –227 (2015). https://doi.org/10.1038/nphoton.2015.29 NPAHBY 1749-4885 Google Scholar

110. 

J. Yi and X. Li, “Estimation of oxygen saturation from erythrocytes by high-resolution spectroscopic optical coherence tomography,” Opt. Lett., 35 (12), 2094 (2010). https://doi.org/10.1364/OL.35.002094 OPLEDP 0146-9592 Google Scholar

111. 

F. E. Robles, S. Chowdhury and A. Wax, “Assessing hemoglobin concentration using spectroscopic optical coherence tomography for feasibility of tissue diagnostics,” Biomed. Opt. Express, 1 (1), 310 –317 (2010). https://doi.org/10.1364/BOE.1.000310 BOEICL 2156-7085 Google Scholar

112. 

J. A. Winkelmann et al., “Spectral contrast optical coherence tomography angiography enables single-scan vessel imaging,” Light Sci. Appl., 8 (1), 7 (2019). https://doi.org/10.1038/s41377-018-0117-7 Google Scholar

113. 

S. Pi et al., “Retinal capillary oximetry with visible light optical coherence tomography,” Proc. Natl. Acad. Sci. U. S. A., 117 (21), 11658 –11666 (2020). https://doi.org/10.1073/pnas.1918546117 Google Scholar

114. 

S. Chen et al., “Retinal oximetry in humans using visible-light optical coherence tomography [Invited],” Biomed. Opt. Express, 8 (3), 1415 (2017). https://doi.org/10.1364/BOE.8.001415 BOEICL 2156-7085 Google Scholar

115. 

W. Song et al., “Visible light optical coherence tomography angiography (vis-OCTA) facilitates local microvascular oximetry in the human retina,” Biomed. Opt. Express, 11 (7), 4037 (2020). https://doi.org/10.1364/BOE.395843 BOEICL 2156-7085 Google Scholar

116. 

D. Huang et al., “Optical coherence tomography,” Science, 254 (5035), 1178 –1181 (1991). https://doi.org/10.1126/science.1957169 SCIEAS 0036-8075 Google Scholar

117. 

J. Yi et al., “Visible-light optical coherence tomography for retinal oximetry,” Opt. Lett., 38 (11), 1796 –1798 (2013). https://doi.org/10.1364/OL.38.001796 OPLEDP 0146-9592 Google Scholar

118. 

S. Chen et al., “Imaging hemodynamic response after ischemic stroke in mouse cortex using visible-light optical coherence tomography,” Biomed. Opt. Express, 7 (9), 3377 (2016). https://doi.org/10.1364/BOE.7.003377 BOEICL 2156-7085 Google Scholar

119. 

P. L. Nesper et al., “OCT angiography and visible-light OCT in diabetic retinopathy,” Vision Res., 139 191 –203 (2017). https://doi.org/10.1016/j.visres.2017.05.006 VISRAM 0042-6989 Google Scholar

120. 

S. Pi et al., “Automated spectroscopic retinal oximetry with visible-light optical coherence tomography,” Biomed. Opt. Express, 9 (5), 2056 (2018). https://doi.org/10.1364/BOE.9.002056 BOEICL 2156-7085 Google Scholar

121. 

B. T. Soetikno et al., “Inner retinal oxygen metabolism in the 50/10 oxygen-induced retinopathy model,” Sci. Rep., 5 16752 (2015). https://doi.org/10.1038/srep16752 SRCEC3 2045-2322 Google Scholar

122. 

W. Liu et al., “Increased retinal oxygen metabolism precedes microvascular alterations in type 1 diabetic mice,” Investig. Ophthalmol. Vis. Sci., 58 (2), 981 –989 (2017). https://doi.org/10.1167/iovs.16-20600 IOVSDA 0146-0404 Google Scholar

123. 

D. J. Cuccia et al., “Quantitation and mapping of tissue optical properties using modulated imaging,” J. Biomed. Opt., 14 (2), 024012 (2009). https://doi.org/10.1117/1.3088140 JBOPFO 1083-3668 Google Scholar

124. 

N. C. Biswal, Y. Xu and Q. Zhu, “Imaging tumor oxyhemoglobin and deoxyhemoglobin concentrations with ultrasound-guided diffuse optical tomography,” Technol. Cancer Res. Treat., 10 (5), 417 –429 (2011). https://doi.org/10.7785/tcrt.2012.500219 Google Scholar

125. 

D. Orive-Miguel et al., “Real-time dual-wavelength time-resolved diffuse optical tomography system for functional brain imaging based on probe-hosted silicon photomultipliers,” Sensors (Switzerland), 20 (10), 2815 (2020). https://doi.org/10.3390/s20102815 Google Scholar

126. 

Y. Wang et al., “Combined diffuse optical tomography and photoacoustic tomography for enhanced functional imaging of small animals: a methodological study on phantoms,” Appl. Opt., 56 (2), 303 (2017). https://doi.org/10.1364/AO.56.000303 APOPAI 0003-6935 Google Scholar

127. 

Y. Hoshi and Y. Yamada, “Overview of diffuse optical tomography and its clinical applications,” J. Biomed. Opt., 21 (9), 091312 (2016). https://doi.org/10.1117/1.JBO.21.9.091312 JBOPFO 1083-3668 Google Scholar

128. 

Y. Shang, T. Li and G. Yu, “Clinical applications of near-infrared diffuse correlation spectroscopy and tomography for tissue blood flow monitoring and imaging,” Physiol. Meas., 38 (4), R1 –R26 (2017). https://doi.org/10.1088/1361-6579/aa60b7 PMEAE3 0967-3334 Google Scholar

129. 

F. Picot et al., “Interstitial imaging with multiple diffusive reflectance spectroscopy projections for in vivo blood vessels detection during brain needle biopsy procedures,” Neurophotonics, 6 (2), 025003 (2019). https://doi.org/10.1117/1.NPh.6.2.025003 Google Scholar

130. 

V. C. Kavuri et al., “Sparsity enhanced spatial resolution and depth localization in diffuse optical tomography,” Biomed. Opt. Express, 3 (5), 943 –957 (2012). https://doi.org/10.1364/BOE.3.000943 BOEICL 2156-7085 Google Scholar

131. 

R. Baikejiang, W. Zhang and C. Li, “Diffuse optical tomography for breast cancer imaging guided by computed tomography: a feasibility study,” J. Xray. Sci. Technol., 25 (3), 341 –355 (2017). https://doi.org/10.3233/XST-16183 Google Scholar

132. 

S. Brigadoi et al., “A 4D neonatal head model for diffuse optical imaging of pre-term to term infants,” Neuroimage, 100 385 –394 (2014). https://doi.org/10.1016/j.neuroimage.2014.06.028 NEIMEF 1053-8119 Google Scholar

133. 

M. A. Khalil et al., “Dynamic diffuse optical tomography imaging of peripheral arterial disease,” Biomed. Opt. Express, 3 (9), 2288 (2012). https://doi.org/10.1364/BOE.3.002288 BOEICL 2156-7085 Google Scholar

134. 

D. A. Boas et al., “Establishing the diffuse correlation spectroscopy signal relationship with blood flow,” Neurophotonics, 3 (3), 031412 (2016). https://doi.org/10.1117/1.NPh.3.3.031412 Google Scholar

135. 

P. D. Mannheimer, “The light-tissue interaction of pulse oximetry,” Anesth. Analg., 105 (6), S10 –S17 (2007). https://doi.org/10.1213/01.ane.0000269522.84942.54 Google Scholar

136. 

J. T. B. Moyle, “Uses and abuses of pulse oximetry,” Arch. Dis. Child., 74 (1), 77 –80 (1996). https://doi.org/10.1136/adc.74.1.77 ADCHAK 0003-9888 Google Scholar

137. 

A. M. Luks and E. R. Swenson, “Pulse oximetry for monitoring patients with COVID-19 at home potential pitfalls and practical guidance,” Ann. Am. Thorac. Soc., 17 (9), 1040 –1046 (2020). https://doi.org/10.1513/AnnalsATS.202005-418FR Google Scholar

138. 

A. Jubran, “Pulse oximetry,” Intensive Care Med., 31 (11), 1598 (2005). https://doi.org/10.1007/s00134-005-2798-7 ICMED9 0342-4642 Google Scholar

139. 

M. Kluckow, “Barriers to the implementation of newborn pulse oximetry screening: a different perspective,” Int. J. Neonatal Screen., 4 (1), 4 (2018). https://doi.org/10.3390/ijns4010004 Google Scholar

140. 

Z. Mosayebi et al., “Evaluation of pulse oximetry in the early diagnosis of cardiac and noncardiac diseases in healthy newborns,” Iran. J. Neonatol., 11 (1), 43 –50 (2020). https://doi.org/10.22038/ijn.2019.38511.1608 Google Scholar

141. 

A. Fawzy et al., “Racial and ethnic discrepancy in pulse oximetry and delayed identification of treatment eligibility among patients with COVID-19,” JAMA Intern Med., 182 (7), 730 –738 (2022). https://doi.org/10.1001/jamainternmed.2022.1906 Google Scholar

142. 

M. W. Sjoding et al., “Racial bias in pulse oximetry measurement,” N. Engl. J. Med., 383 (25), 2477 –2478 (2020). https://doi.org/10.1056/NEJMc2029240 NEJMAG 0028-4793 Google Scholar

143. 

A. Jubran, “Pulse oximetry,” Intensive Care Med., 19 (1), 1 –7 (2005). https://doi.org/10.1186/s13054-015-0984-8 ICMED9 0342-4642 Google Scholar

144. 

E. D. Chan, M. M. Chan, M. M. Chan, “Pulse oximetry: understanding its basic principles facilitates appreciation of its limitations,” Respir. Med., 107 (6), 789 –799 (2013). https://doi.org/10.1016/j.rmed.2013.02.004 RMEDEY 0954-6111 Google Scholar

145. 

R. W. C. G. R. Wijshoff et al., “Reducing motion artifacts in photoplethysmograms by using relative sensor motion: phantom study,” J. Biomed. Opt., 17 (11), 117007 (2012). https://doi.org/10.1117/1.JBO.17.11.117007 JBOPFO 1083-3668 Google Scholar

146. 

H. Zhang et al., “Wireless, battery-free optoelectronic systems as subdermal implants for local tissue oximetry,” Sci. Adv., 5 (3), (2019). https://doi.org/10.1126/sciadv.aaw0873 STAMCV 1468-6996 Google Scholar

147. 

J. Nixdorff et al., “Comparison of transmittance and reflectance pulse oximetry in anesthetized dogs,” Front. Vet. Sci., 8 (April), 1 –7 (2021). https://doi.org/10.3389/fvets.2021.643966 Google Scholar

148. 

L. E. Mackenzie and A. R. Harvey, “Oximetry using multispectral imaging: theory and application,” J. Opt. (United Kingdom), 20 (6), 063501 (2018). https://doi.org/10.1088/2040-8986/aab74c Google Scholar

149. 

M. Nitzan and S. Engelberg, “Three-wavelength technique for the measurement of oxygen saturation in arterial blood and in venous blood,” J. Biomed. Opt., 14 (2), 024046 (2009). https://doi.org/10.1117/1.3120496 JBOPFO 1083-3668 Google Scholar

150. 

Q. J. W. Milner and G. R. Mathews, “An assessment of the accuracy of pulse oximeters,” Anaesthesia, 67 (4), 396 –401 (2012). https://doi.org/10.1111/j.1365-2044.2011.07021.x Google Scholar

151. 

M. Oura et al., “Calibration system for pulse spectrophotometry using a double-layer pulsation flow-cell,” in Proc. 31st Annu. Int. Conf. IEEE Eng. Med. Biol. Soc. Eng. Futur. Biomed. EMBC, 896 –899 (2009). https://doi.org/10.1109/IEMBS.2009.5334889 Google Scholar

152. 

V. N. D. Le et al., “Calibration of spectral imaging devices with oxygenation-controlled phantoms: introducing a simple gel-based hemoglobin model,” Front. Phys., 7 (November), 1 –7 (2019). https://doi.org/10.3389/fphy.2019.00192 Google Scholar

153. 

J. E. Yount, “Devices and procedures for in vitro testing of pulse oximetry monitors,” (1990). Google Scholar

154. 

S. J. Barker and K. K. Tremper, “The effect of carbon monoxide inhalation on pulse oximetry and transcutaneous PO2,” Anesthesiology, 66 (5), 677 –679 (1987). https://doi.org/10.1097/00000542-198705000-00014 ANESAV 0003-3022 Google Scholar

155. 

Clinical Dynamics Corporation, “Smart Sat TM Pulse Oximetry Analyzer,” 1 –42 (2004). Google Scholar

156. 

M. Nitzan et al., “Calibration-free pulse oximetry based on two wavelengths in the infrared - a preliminary study,” Sensors (Switzerland), 14 (4), 7420 –7434 (2014). https://doi.org/10.3390/s140407420 Google Scholar

157. 

S. J. Barker et al., “Measurement of carboxyhemoglobin and methemoglobin by pulse oximetry: a human volunteer study,” Anesthesiology, 105 (5), 892 –897 (2006). https://doi.org/10.1097/00000542-200611000-00008 ANESAV 0003-3022 Google Scholar

158. 

P. E. Bicikler, J. R. Feiner and J. W. Severinghaus, “Effects of skin pigmentation on pulse oximeter accuracy at low saturation,” Anesthesiology, 102 (4), 715 –719 (2005). https://doi.org/10.1097/00000542-200504000-00004 ANESAV 0003-3022 Google Scholar

159. 

Q. Li et al., “Review of spectral imaging technology in biomedical engineering: achievements and challenges,” J. Biomed. Opt., 18 (10), 100901 (2013). https://doi.org/10.1117/1.JBO.18.10.100901 JBOPFO 1083-3668 Google Scholar

160. 

P. Lukes et al., “Narrow band imaging (NBI) — endoscopic method for detection of head and neck cancer,” Endoscopy, 5 75 –87 (2013). https://doi.org/10.5772/52738 ENDCAM Google Scholar

161. 

Q. Li et al., “Estimation of tissue oxygen saturation from RGB images and sparse hyperspectral signals based on conditional generative adversarial network,” Int. J. Comput. Assist. Radiol. Surg., 14 (6), 987 –995 (2019). https://doi.org/10.1007/s11548-019-01940-2 Google Scholar

162. 

D. S. Terman et al., “Sickle erythrocytes target cytotoxics to hypoxic tumor microvessels and potentiate a tumoricidal response,” PLoS One, 8 (1), e52543 (2013). https://doi.org/10.1371/journal.pone.0052543 POLNCL 1932-6203 Google Scholar

163. 

H. L. Offerhaus, S. E. Bohndiek and A. R. Harvey, “Hyperspectral imaging in biomedical applications,” J. Opt. (United Kingdom), 21 (1), 010202 (2019). https://doi.org/10.1088/2040-8986/aaf2a0 Google Scholar

164. 

E. Häggblad et al., “Reflection spectroscopy of analgesized skin,” Microvasc. Res., 62 (3), 392 –400 (2001). https://doi.org/10.1006/mvre.2001.2358 MIVRA6 0026-2862 Google Scholar

165. 

L. Giannoni, F. Lange and I. Tachtsidis, “Hyperspectral imaging solutions for brain tissue metabolic and hemodynamic monitoring: past, current and future developments,” J. Opt. (United Kingdom), 20 (4), 44009 (2018). https://doi.org/10.1088/2040-8986/aab3a6 Google Scholar

166. 

T. W. Sawyer et al., “Opti-MSFA: a toolbox for generalized design and optimization of multispectral filter arrays,” Opt. Express, 30 (5), 7591 (2022). https://doi.org/10.1364/OE.446767 OPEXFF 1094-4087 Google Scholar

167. 

M. Taylor-Williams et al., “Spectrally tailored ‘hyperpixel’ filter arrays for imaging of chemical compositions,” Proc. SPIE, 11954 1195406 (2022). https://doi.org/10.1117/12.2606917 PSISDG 0277-786X Google Scholar

168. 

S. Beg, A. Wilson and K. Ragunath, “The use of optical imaging techniques in the gastrointestinal tract,” Frontline Gastroenterol., 7 (3), 207 –215 (2016). https://doi.org/10.1136/flgastro-2015-100563 Google Scholar

169. 

S. J. Spechler et al., “American gastroenterological association technical review on the management of Barrett’s esophagus,” Gastroenterology, 140 (3), e18 –e52 (2011). https://doi.org/10.1053/j.gastro.2011.01.031 GASTAB 0016-5085 Google Scholar

170. 

J. Yoon et al., “First experience in clinical application of hyperspectral endoscopy for evaluation of colonic polyps,” J. Biophotonics, 14 (9), 1 –9 (2021). https://doi.org/10.1002/jbio.202100078 Google Scholar

171. 

D. J. Waterhouse et al., “Spectral endoscopy enhances contrast for neoplasia in surveillance of Barrett’s esophagus,” Cancer Res., 81 (12), 3415 –3425 (2021). https://doi.org/10.1158/0008-5472.CAN-21-0474 CNREA8 0008-5472 Google Scholar

172. 

V. Dremin et al., “Dynamic evaluation of blood flow microcirculation by combined use of the laser Doppler flowmetry and high-speed videocapillaroscopy methods,” J. Biophotonics, 12 (6), 1 –7 (2019). https://doi.org/10.1002/jbio.201800317 Google Scholar

173. 

M. V. Volkov et al., “Evaluation of blood microcirculation parameters by combined use of laser Doppler flowmetry and videocapillaroscopy methods,” Proc. SPIE, 10336 1033607 (2017). https://doi.org/10.1117/12.2267955 PSISDG 0277-786X Google Scholar

174. 

U. Baran, L. Shi and R. K. Wang, “Capillary blood flow imaging within human finger cuticle using optical microangiography,” J. Biophotonics, 8 (1–2), 46 –51 (2015). https://doi.org/10.1002/jbio.201300154 Google Scholar

175. 

W. F. Yip et al., “Reliability and determinants of retinal vessel oximetry measurements in healthy eyes,” Invest. Ophthalmol. Vis. Sci., 55 (11), 7104 –7110 (2014). https://doi.org/10.1167/iovs.13-13854 IOVSDA 0146-0404 Google Scholar

176. 

A. K. Garg et al., “Advances in retinal oximetry,” Transl. Vis. Sci. Technol., 10 (2), 5 –18 (2021). https://doi.org/10.1167/tvst.10.2.5 Google Scholar

177. 

J. M. Kainerstorfer et al., “Quantitative principal component model for skin chromophore mapping using multi-spectral images and spatial priors,” Biomed. Opt. Express, 2 (5), 1040 (2011). https://doi.org/10.1364/BOE.2.001040 BOEICL 2156-7085 Google Scholar

178. 

K. Kikuchi, Y. Masuda and T. Hirao, “Imaging of hemoglobin oxygen saturation ratio in the face by spectral camera and its application to evaluate dark circles,” Ski. Res. Technol., 19 (4), 499 –507 (2013). https://doi.org/10.1111/srt.12074 Google Scholar

179. 

Q. He and R. Wang, “Hyperspectral imaging enabled by an unmodified smartphone for analyzing skin morphological features and monitoring hemodynamics,” Biomed. Opt. Express, 11 (2), 895 (2020). https://doi.org/10.1364/BOE.378470 BOEICL 2156-7085 Google Scholar

180. 

E. Zherebtsov et al., “Hyperspectral imaging of human skin aided by artificial neural networks,” Biomed. Opt. Express, 10 (7), 3545 (2019). https://doi.org/10.1364/BOE.10.003545 BOEICL 2156-7085 Google Scholar

181. 

N. T. Clancy et al., “Intraoperative measurement of bowel oxygen saturation using a multispectral imaging laparoscope,” Biomed. Opt. Express, 6 (10), 4179 (2015). https://doi.org/10.1364/BOE.6.004179 BOEICL 2156-7085 Google Scholar

182. 

T. Pruimboom et al., “Perioperative hyperspectral imaging to assess mastectomy skin flap and DIEP flap perfusion in immediate autologous breast reconstruction: a pilot study,” Diagnostics, 12 (1), 184 (2022). https://doi.org/10.3390/diagnostics12010184 Google Scholar

183. 

J. M. Eichenholz et al., “Real-time megapixel multispectral bioimaging,” Proc. SPIE, 7568 75681L (2010). https://doi.org/10.1117/12.842563 PSISDG 0277-786X Google Scholar

184. 

R. Calvini, A. Ulrici and J. M. Amigo, “Growing applications of hyperspectral and multispectral imaging,” Data Handl. Sci. Technol., 32 605 –629 (2020). https://doi.org/10.1016/B978-0-444-63977-6.00024-9 DHSTEV Google Scholar

185. 

T. W. Sawyer, C. Williams and S. E. Bohndiek, “Spectral band selection and tolerancing for multispectral filter arrays,” in Front. Opt. - Proc. Front. Opt. + Laser Sci. APS/DLS, 3 –5 (2019). https://doi.org/10.1364/FIO.2019.JW4A.126 Google Scholar

186. 

C. Williams et al., “Grayscale-to-color: scalable fabrication of custom multispectral filter arrays,” ACS Photonics, 6 (12), 3132 –3141 (2019). https://doi.org/10.1021/acsphotonics.9b01196 Google Scholar

187. 

R. Wu et al., “Optimized multi-spectral filter arrays for spectral reconstruction,” Sensors (Switzerland), 19 (13), 2905 (2019). https://doi.org/10.3390/s19132905 Google Scholar

188. 

S. Gioux, A. Mazhar and D. J. Cuccia, “Spatial frequency domain imaging in 2019: principles, applications, and perspectives,” J. Biomed. Opt., 24 (7), 071613 (2019). https://doi.org/10.1117/1 JBOPFO 1083-3668 Google Scholar

189. 

R. H. Wilson et al., “High-speed spatial frequency domain imaging of rat cortex detects dynamic optical and physiological properties following cardiac arrest and resuscitation,” Neurophotonics, 4 (4), 045008 (2017). https://doi.org/10.1117/1.NPh.4.4.045008 Google Scholar

190. 

M. T. Ghijsen et al., “Quantitative real-time optical imaging of the tissue metabolic rate of oxygen consumption,” J. Biomed. Opt., 23 (3), 036013 (2018). https://doi.org/10.1117/1.JBO.23.3.036013 JBOPFO 1083-3668 Google Scholar

191. 

M. Schmidt et al., “Real-time, wide-field, and quantitative oxygenation imaging using spatiotemporal modulation of light,” J. Biomed. Opt., 24 (7), 071610 (2019). https://doi.org/10.1117/1.JBO.24.7.071610 JBOPFO 1083-3668 Google Scholar

192. 

C. Weinkauf et al., “Near-instant noninvasive optical imaging of tissue perfusion for vascular assessment,” J. Vasc. Surg., 69 (2), 555 –562 (2019). https://doi.org/10.1016/j.jvs.2018.06.202 Google Scholar

193. 

S. Jett et al., “Stratification of microvascular disease severity in the foot using spatial frequency domain imaging,” J. Diabetes Sci. Technol., 19322968211024666 (2021). https://doi.org/10.1177/19322968211024666 Google Scholar

194. 

A. Ponticorvo et al., “Spatial frequency domain imaging (SFDI) of clinical burns: a case report,” Burn. Open, 4 (2), 67 –71 (2020). https://doi.org/10.1016/j.burnso.2020.02.004 Google Scholar

195. 

J. Sun et al., “Enhancing in vivo tumor boundary delineation with structured illumination fluorescence molecular imaging and spatial gradient mapping,” J. Biomed. Opt., 21 (8), 080502 (2016). https://doi.org/10.1117/1.JBO.21.8.080502 JBOPFO 1083-3668 Google Scholar

196. 

D. Wirth et al., “Feasibility of using spatial frequency-domain imaging intraoperatively during tumor resection,” J. Biomed. Opt., 24 (7), 071608 (2018). https://doi.org/10.1117/1.JBO.24.7.071608 JBOPFO 1083-3668 Google Scholar

197. 

P. A. Valdes et al., “qF-SSOP: real-time optical property corrected fluorescence imaging,” Biomed. Opt. Express, 8 (8), 3597 (2017). https://doi.org/10.1364/BOE.8.003597 BOEICL 2156-7085 Google Scholar

198. 

Y. Garini, I. T. Young and G. McNamara, “Spectral imaging: principles and applications,” Cytom. Part A, 69 (8), 735 –747 (2006). https://doi.org/10.1002/cyto.a.20311 1552-4922 Google Scholar

199. 

J. Wei and X. Wang, “An overview on linear unmixing of hyperspectral data,” Math. Probl. Eng., 2020 3735403 (2020). https://doi.org/10.1155/2020/3735403 Google Scholar

200. 

M. D. Mura, J. Chanussot and A. Plaza, An Overview on Hyperspectral Unmixing, GIPSA-Lab, Grenoble Inst. Technology(2014). Google Scholar

201. 

J. S. Bhatt and M. V. Joshi, “Deep learning in hyperspectral unmixing: a review,” in Int. Geosci. Remote Sens. Symp., 2189 –2192 (2020). https://doi.org/10.1109/IGARSS39084.2020.9324546 Google Scholar

202. 

A. Grigoroiu, J. Yoon and S. E. Bohndiek, “Deep learning applied to hyperspectral endoscopy for online spectral classification,” Sci. Rep., 10 3947 (2020). https://doi.org/10.1038/s41598-020-60574-6 SRCEC3 2045-2322 Google Scholar

203. 

S. Li et al., “Deep learning for hyperspectral image classification: an overview,” IEEE Trans. Geosci. Remote Sens., 57 (9), 6690 –6709 (2019). https://doi.org/10.1109/TGRS.2019.2907932 IGRSD2 0196-2892 Google Scholar

204. 

J. Gröhl et al., “Learned spectral decoloring enables photoacoustic oximetry,” Sci. Rep., 11 6565 (2021). https://doi.org/10.1038/s41598-021-83405-8 SRCEC3 2045-2322 Google Scholar

205. 

S. Agrawal et al., “Functional, molecular and structural imaging using LED-based photoacoustic and ultrasound imaging system,” Proc. SPIE, 11240 112405A (2020). https://doi.org/10.1117/12.2547048 PSISDG 0277-786X Google Scholar

206. 

D. Lighter et al., “Multispectral, non-contact diffuse optical tomography of healthy human finger joints,” Biomed. Opt. Express, 9 (4), 1445 (2018). https://doi.org/10.1364/BOE.9.001445 BOEICL 2156-7085 Google Scholar

207. 

S. P. Chong et al., “Cerebral metabolic rate of oxygen (CMRO_2) assessed by combined Doppler and spectroscopic OCT,” Biomed. Opt. Express, 6 (10), 3941 (2015). https://doi.org/10.1364/BOE.6.003941 BOEICL 2156-7085 Google Scholar

208. 

E. Brown, J. Brunker and S. E. Bohndiek, “Photoacoustic imaging as a tool to probe the tumour microenvironment,” DMM Dis. Model. Mech., 12 (7), dmm039636 (2019). https://doi.org/10.1242/dmm.039636 Google Scholar

209. 

I. Steinberg et al., “Photoacoustic clinical imaging,” Photoacoustics, 14 77 –98 (2019). https://doi.org/10.1016/j.pacs.2019.05.001 Google Scholar

210. 

T. Shiina, M. Toi and T. Yagi, “Development and clinical translation of photoacoustic mammography,” Biomed. Eng. Lett., 8 (2), 157 –165 (2018). https://doi.org/10.1007/s13534-018-0070-7 Google Scholar

211. 

P. G. Anderson et al., “Broadband optical mammography: chromophore concentration and hemoglobin saturation contrast in breast cancer,” PLoS One, 10 (3), 1 –23 (2015). https://doi.org/10.1371/journal.pone.0117322 POLNCL 1932-6203 Google Scholar

212. 

L. He et al., “Noncontact diffuse correlation tomography of human breast tumor,” J. Biomed. Opt., 20 (8), 086003 (2015). https://doi.org/10.1117/1.JBO.20.8.086003 JBOPFO 1083-3668 Google Scholar

213. 

E. I. Neuschler et al., “Breast imaging: optoacoustic imaging to diagnose benign and malignant breast masses Neuschler et al materials and methods,” Radiology, 287 (2), (2017). https://doi.org/10.1148/radiol.2017172228 RADLAX 0033-8419 Google Scholar

214. 

M. Omar, J. Aguirre and V. Ntziachristos, “Optoacoustic mesoscopy for biomedicine,” Nat. Biomed. Eng., 3 (5), 354 –370 (2019). https://doi.org/10.1038/s41551-019-0377-4 Google Scholar

215. 

G. L. G. Menezes et al., “Downgrading of breast masses suspicious for cancer by using optoacoustic breast imaging,” Radiology, 288 (2), 355 –365 (2018). https://doi.org/10.1148/radiol.2018170500 RADLAX 0033-8419 Google Scholar

216. 

W. Roll et al., “Multispectral optoacoustic tomography of benign and malignant thyroid disorders: a pilot study,” J. Nucl. Med., 60 (10), 1461 –1466 (2019). https://doi.org/10.2967/jnumed.118.222174 JNMEAQ 0161-5505 Google Scholar

217. 

S. Zackrisson, S. M. W. Y. Van de Ven and S. S. Gambhir, “Light in and sound out: Emerging translational strategies for photoacoustic imaging,” Cancer Res., 74 (4), 979 –1004 (2014). https://doi.org/10.1158/0008-5472.CAN-13-2387 CNREA8 0008-5472 Google Scholar

218. 

F. Knieling et al., “Multispectral optoacoustic tomography for assessment of Crohn’s disease activity,” New England J. Med., 376 (13), 1292 –1294 (2017). https://doi.org/10.1056/NEJMc1612455 NEJMBH Google Scholar

219. 

M. Erfanzadeh and Q. Zhu, “Photoacoustic imaging with low-cost sources; a review,” Photoacoustics, 14 1 –11 (2019). https://doi.org/10.1016/j.pacs.2019.01.004 Google Scholar

220. 

D. Das et al., “Another decade of photoacoustic imaging,” Phys. Med. Biol., 66 (5), 05TR01 (2021). https://doi.org/10.1088/1361-6560/abd669 PHMBA7 0031-9155 Google Scholar

221. 

A. B. E. Attia et al., “A review of clinical photoacoustic imaging: current and future trends,” Photoacoustics, 16 100144 (2019). https://doi.org/10.1016/j.pacs.2019.100144 Google Scholar

222. 

B. T. Cox, J. G. Laufer and P. C. Beard, “Quantitative photoacoustic image reconstruction using fluence dependent chromophores,” Biomed. Opt. Express, 1 (1), 201 (2010). https://doi.org/10.1364/BOE.1.000201 BOEICL 2156-7085 Google Scholar

223. 

H. Singh et al., “Mapping cortical haemodynamics during neonatal seizures using diffuse optical tomography: a case study,” NeuroImage Clin., 5 256 –265 (2014). https://doi.org/10.1016/j.nicl.2014.06.012 Google Scholar

224. 

M. D. Wheelock, J. P. Culver and A. T. Eggebrecht, “High-density diffuse optical tomography for imaging human brain function,” Rev. Sci. Instrum., 90 (5), 051101 (2019). https://doi.org/10.1063/1.5086809 RSINAK 0034-6748 Google Scholar

225. 

A. P. Gibson, J. C. Hebden and S. R. Arridge, “Recent advances in diffuse optical imaging,” Phys. Med. Biol., 50 (4), R1 –R43 (2005). https://doi.org/10.1088/0031-9155/50/4/R01 PHMBA7 0031-9155 Google Scholar

226. 

W. Zhi et al., “Predicting treatment response of breast cancer to neoadjuvant chemotherapy using ultrasound-guided diffuse optical tomography,” Transl. Oncol., 11 (1), 56 –64 (2018). https://doi.org/10.1016/j.tranon.2017.10.011 Google Scholar

227. 

R. Choe and T. Durduran, “Diffuse optical monitoring of the neoadjuvant breast cancer therapy,” IEEE J. Sel. Top. Quantum Electron., 18 (4), 1367 –1386 (2012). https://doi.org/10.1109/JSTQE.2011.2177963 IJSQEN 1077-260X Google Scholar

228. 

F. Scholkmann et al., “A review on continuous wave functional near-infrared spectroscopy and imaging instrumentation and methodology,” Neuroimage, 85 6 –27 (2014). https://doi.org/10.1016/j.neuroimage.2013.05.004 NEIMEF 1053-8119 Google Scholar

229. 

K. A. S. Mithun and W. Xia, “Portable and affordable light source-based photoacoustic tomography,” Sensors, 20 (21), 6173 (2020). https://doi.org/10.3390/s20216173 SNSRES 0746-9462 Google Scholar

230. 

V. V. Beschastnov et al., “Current methods for the assessment of oxygen status and biotissue microcirculation condition: diffuse optical spectroscopy (review),” Sovrem. Tehnol. v Med., 10 (4), 183 –194 (2018). https://doi.org/10.17691/stm2018.10.4.22 Google Scholar

231. 

A. Curtin et al., “A systematic review of integrated functional near-infrared spectroscopy (fNIRS) and transcranial magnetic stimulation (TMS) studies,” Front. Neurosci., 13 84 (2019). https://doi.org/10.3389/fnins.2019.00084 1662-453X Google Scholar

232. 

P. Pinti et al., “The present and future use of functional near-infrared spectroscopy (fNIRS) for cognitive neuroscience,” Ann. N. Y. Acad. Sci., 1464 5 –29 (2018). https://doi.org/10.1111/nyas.13948 ANYAA9 0077-8923 Google Scholar

233. 

A. Pifferi et al., “New frontiers in time-domain diffuse optics, a review,” J. Biomed. Opt., 21 (9), 091310 (2016). https://doi.org/10.1117/1.JBO.21.9.091310 JBOPFO 1083-3668 Google Scholar

234. 

E. E. Vidal-Rosas et al., “Evaluating a new generation of wearable high-density diffuse optical tomography technology via retinotopic mapping of the adult visual cortex,” Neurophotonics, 8 (2), 025002 (2021). https://doi.org/10.1117/1.NPh.8.2.025002 Google Scholar

235. 

G. Di Leo et al., “Optical imaging of the breast: basic principles and clinical applications,” Am. J. Roentgenol., 209 (1), 230 –238 (2017). https://doi.org/10.2214/AJR.16.17220 AJROAM 0092-5381 Google Scholar

236. 

F. Y. Wong et al., “Impaired autoregulation in preterm infants identified by using spatially resolved spectroscopy,” Pediatrics, 121 (3), e604 –e611 (2008). https://doi.org/10.1542/peds.2007-1487 PEDIAU 0031-4005 Google Scholar

237. 

H. S. Yazdi et al., “Mapping breast cancer blood flow index, composition, and metabolism in a human subject using combined diffuse optical spectroscopic imaging and diffuse correlation spectroscopy,” J. Biomed. Opt., 22 (4), 045003 (2017). https://doi.org/10.1117/1.JBO.22.4.045003 JBOPFO 1083-3668 Google Scholar

238. 

I. Fredriksson and M. Larsson, “On the equivalence and differences between laser Doppler flowmetry and laser speckle contrast analysis,” J. Biomed. Opt., 21 (12), 126018 (2016). https://doi.org/10.1117/1.JBO.21.12.126018 JBOPFO 1083-3668 Google Scholar

239. 

H. Jonasson et al., “Oxygen saturation, red blood cell tissue fraction and speed resolved perfusion - a new optical method for microcirculatory assessment,” Microvasc. Res., 102 70 –77 (2015). https://doi.org/10.1016/j.mvr.2015.08.006 MIVRA6 0026-2862 Google Scholar

240. 

W. Heeman et al., “Clinical applications of laser speckle contrast imaging: a review,” J. Biomed. Opt., 24 (08), 1 (2019). https://doi.org/10.1117/1.JBO.24.8.080901 JBOPFO 1083-3668 Google Scholar

241. 

Q. Fang and S. Yan, “MCX Cloud—a modern, scalable, high-performance and in-browser Monte Carlo simulation platform with cloud computing,” J. Biomed. Opt., 27 (8), 083008 (2022). https://doi.org/10.1117/1.JBO.27.8.083008 JBOPFO 1083-3668 Google Scholar

242. 

S. R. P. K. Lanka et al., “Multi-laboratory performance assessment of diffuse optics instruments: the BitMap exercise,” J. Biomed. Opt., 27 (7), 074716 (2022). https://doi.org/10.1117/1.JBO.27.7.074716 Google Scholar

243. 

L. Di Sieno et al., “Solid heterogeneous phantoms for multimodal ultrasound and diffuse optical imaging: an outcome of the SOLUS project for standardization,” Proc. SPIE, 11075 1107516 (2019). https://doi.org/10.1117/12.2526645 PSISDG 0277-786X Google Scholar

244. 

L. Kagemann et al., “Spectral oximetry assessed with high-speed ultra-high- resolution optical coherence tomography Larry,” J. Biomed. Opt., 12 (4), 041212 (2007). https://doi.org/10.1117/1.2772655 JBOPFO 1083-3668 Google Scholar

245. 

M. J. Casper et al., “Capillary refill—the key to assessing dermal capillary capacity and pathology in optical coherence tomography angiography,” Lasers Surg. Med., 52 (7), 653 –658 (2020). https://doi.org/10.1002/lsm.23188 LSMEDI 0196-8092 Google Scholar

246. 

A. H. Kashani et al., “Optical coherence tomography angiography: A comprehensive review of current methods and clinical applications,” Prog. Retin. Eye Res., 60 66 –100 (2017). https://doi.org/10.1016/j.preteyeres.2017.07.002 PRTRES 1350-9462 Google Scholar

247. 

M. Liu and W. Drexler, “Optical coherence tomography angiography and photoacoustic imaging in dermatology,” Photochem. Photobiol. Sci., 18 (5), 945 –962 (2019). https://doi.org/10.1039/C8PP00471D PPSHCB 1474-905X Google Scholar

248. 

R. Leitgeb, C. K. Hitzenberger and A. F. Fercher, “Performance of fourier domain vs. time domain optical coherence tomography,” Opt. Express, 11 (8), 889 –894 (2003). https://doi.org/10.1364/OE.11.000889 OPEXFF 1094-4087 Google Scholar

249. 

B. Povazay et al., “Visible light optical coherence tomography,” J. Biomed. Opt., 22 (12), 121707 (2002). https://doi.org/10.1117/12.470466 JBOPFO 1083-3668 Google Scholar

250. 

S. P. Chong et al., “Quantitative microvascular hemoglobin mapping using visible light spectroscopic optical coherence tomography,” Biomed. Opt. Express, 6 (4), 1429 (2015). https://doi.org/10.1364/BOE.6.001429 BOEICL 2156-7085 Google Scholar

251. 

S. Chen, J. Yi and H. F. Zhang, “Measuring oxygen saturation in retinal and choroidal circulations in rats using visible light optical coherence tomography angiography,” Biomed. Opt. Express, 6 (8), 2840 (2015). https://doi.org/10.1364/BOE.6.002840 BOEICL 2156-7085 Google Scholar

252. 

G. J. Tearney et al., “In vivo endoscopic optical biopsy with optical coherence tomography,” Science, 276 (5321), 2037 –2039 (1997). https://doi.org/10.1126/science.276.5321.2037 SCIEAS 0036-8075 Google Scholar

253. 

B. C. Quirk et al., “In situ imaging of lung alveoli with an optical coherence tomography needle probe,” J. Biomed. Opt., 16 (3), 036009 (2011). https://doi.org/10.1117/1.3556719 JBOPFO 1083-3668 Google Scholar

254. 

H. Yabushita et al., “Characterization of human atherosclerosis by optical coherence tomography,” Circulation, 106 (13), 1640 –1645 (2002). https://doi.org/10.1161/01.CIR.0000029927.92825.F6 CIRCAZ 0009-7322 Google Scholar

255. 

J. A. Evans et al., “Optical coherence tomography to identify intramucosal carcinoma and high-grade dysplasia in Barrett’s esophagus,” Clin. Gastroenterol. Hepatol., 4 (1), 38 –43 (2006). https://doi.org/10.1016/S1542-3565(05)00746-9 Google Scholar

256. 

D. Lorenser, R. A. Mclaughlin, D. D. Sampson, “Optical Coherence Tomography in a Needle Format,” Optical Coherence Tomography, 2015th ed.Springer, Switzerland (2015). Google Scholar

257. 

S. Song, J. Xu and R. K. Wang, “Long-range and wide field of view optical coherence tomography for in vivo 3D imaging of large volume object based on akinetic programmable swept source,” Biomed. Opt. Express, 7 (11), 4734 (2016). https://doi.org/10.1364/BOE.7.004734 BOEICL 2156-7085 Google Scholar

258. 

M. Ulrich et al., “Dynamic optical coherence tomography in dermatology,” Dermatology, 232 (3), 298 –311 (2016). https://doi.org/10.1159/000444706 DERMEI 0742-3217 Google Scholar

259. 

H. Liang et al., “Optical coherence tomography for art conservation and archaeology,” O3A Opt. Arts Archit. Archaeol., 6618 661805 (2007). https://doi.org/10.1117/12.726032 Google Scholar

260. 

X. Qi, M. V. Sivak, A. M. Rollins, “Optical coherence tomography for gastrointestinal endoscopy,” Optical Coherence Tomography Technology and Applications, 1047 –1081 2nd ed.Springer International Publishing, Switzerland (2015). Google Scholar

261. 

G. J. Tearney et al., “Imaging coronary atherosclerosis and vulnerable plaques with optical coherence tomography,” Optical Coherence Tomography Technology and Applications, 2109 –2130 2nd ed.Springer, Berlin, Heidelberg (2015). Google Scholar

262. 

K. Liang et al., “Tethered capsule en face optical coherence tomography for imaging Barrett’s oesophagus in unsedated patients,” BMJ Open Gastroenterol., 7 (1), e000444 (2020). https://doi.org/10.1136/bmjgast-2020-000444 Google Scholar

263. 

Z. Hosseinaee, J. A. Tummon Simmons and P. H. Reza, “Dual-modal photoacoustic imaging and optical coherence tomography [review],” Front. Phys., 8 1 –19 (2021). https://doi.org/10.3389/fphy.2020.616618 Google Scholar

264. 

H. Jonasson et al., “Validation of speed-resolved laser Doppler perfusion in a multimodal optical system using a blood-flow phantom,” J. Biomed. Opt., 24 (9), 095002 (2019). https://doi.org/10.1117/1.JBO.24.9.095002 JBOPFO 1083-3668 Google Scholar

265. 

K. F. Ma et al., “Laser Doppler flowmetry combined with spectroscopy to determine peripheral tissue perfusion and oxygen saturation: a pilot study in healthy volunteers and patients with peripheral arterial disease,” J. Pers. Med., 12 (6), 853 (2022). https://doi.org/10.3390/jpm12060853 Google Scholar

266. 

G. Wang et al., “Impact of local thermal stimulation on the correlation between oxygen saturation and speed-resolved blood perfusion,” Sci. Rep., 10 183 (2020). https://doi.org/10.1038/s41598-019-57067-6 SRCEC3 2045-2322 Google Scholar

267. 

A. Pellicer et al., “The SafeBoosC phase II randomised clinical trial: A treatment guideline for targeted near-infrared-derived cerebral tissue oxygenation versus standard treatment in extremely preterm infants,” Neonatology, 104 (3), 171 –178 (2013). https://doi.org/10.1159/000351346 Google Scholar

Biography

Michaela Taylor-Williams received her BSc (Honours) degree in electrical and electronic engineering science from the University of Western Australia in 2018. She subsequently secured a Sir John Monash Foundation Scholarship and Cambridge Trust Scholarship to undertake her PhD at the University of Cambridge, where she is developing compact biomedical optical imaging systems for imaging tissue vasculature in microscopy and endoscopy configurations.

Graham Spicer received his PhD in chemical engineering from Northwestern University in 2019 and a postdoctoral fellowship at Harvard Medical School focused on advancing the state of the art in optical coherence tomography. He is currently a postdoctoral research fellow at the University of Cambridge, developing new hyperspectral and depth-resolved endoscopic imaging tools for early cancer detection.

Gemma Bale received her PhD in biomedical optics from University College London, pioneering broadband near-infrared spectroscopy as a tool to monitor metabolism in brain injury. Following her postdoctoral research, expanding these efforts to clinical application, she started her laboratory at the University of Cambridge where she is the Gianna Angelopoulos Lecturer in medical therapeutics in the Departments of Physics and Engineering.

Sarah E. Bohndiek received her PhD in radiation physics from University College London in 2008 and then worked in both the UK (at Cambridge) and the United States (at Stanford) as a postdoctoral fellow in molecular imaging. She joined the University of Cambridge in 2013 and is currently a professor of biomedical physics in the Department of Physics, the Cancer Research UK Cambridge Institute. Her team uses multispectral imaging methods to study cancer evolution.

© The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Michaela Taylor-Williams, Graham Spicer, Gemma Bale, and Sarah E. Bohndiek "Noninvasive hemoglobin sensing and imaging: optical tools for disease diagnosis," Journal of Biomedical Optics 27(8), 080901 (3 August 2022). https://doi.org/10.1117/1.JBO.27.8.080901
Received: 6 April 2022; Accepted: 27 June 2022; Published: 3 August 2022
Lens.org Logo
CITATIONS
Cited by 10 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Tissue optics

Oximetry

Optical coherence tomography

Absorption

Imaging systems

Optical imaging

Blood

RELATED CONTENT


Back to Top