Next Article in Journal
Applicability Evaluation of Soil Algae Pipe Assay in Silver Nanoparticle-Contaminated Soils
Previous Article in Journal
AnyGesture: Arbitrary One-Handed Gestures for Augmented, Virtual, and Mixed Reality Applications
Previous Article in Special Issue
Broadband Passively Mode-Locked Fiber Laser with DNA Aqueous Solution as Saturable Absorber
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Optical Modalities for Research, Diagnosis, and Treatment of Stroke and the Consequent Brain Injuries

1
Medical Device Development Center, Daegu-Gyeongbuk Medical Innovation Foundation (K-MEDI hub), Daegu 41061, Korea
2
Department of Medical Science, School of Medicine, Kyungpook National University, Daegu 41944, Korea
3
Preclinical Research Center, Daegu-Gyeongbuk Medical Innovation Foundation (K-MEDI hub), Daegu 41061, Korea
4
Laboratory Animal Center, Osong Medical Innovation Foundation, Cheonju 28160, Korea
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(4), 1891; https://doi.org/10.3390/app12041891
Submission received: 9 December 2021 / Revised: 2 February 2022 / Accepted: 7 February 2022 / Published: 11 February 2022
(This article belongs to the Special Issue Photonics in BioMedical Progress)

Abstract

:
Stroke is the second most common cause of death and third most common cause of disability worldwide. Therefore, it is an important disease from a medical standpoint. For this reason, various studies have developed diagnostic and therapeutic techniques for stroke. Among them, developments and applications of optical modalities are being extensively studied. In this article, we explored three important optical modalities for research, diagnostic, and therapeutics for stroke and the brain injuries related to it: (1) photochemical thrombosis to investigate stroke animal models; (2) optical imaging techniques for in vivo preclinical studies on stroke; and (3) optical neurostimulation based therapy for stroke. We believe that an exploration and an analysis of previous studies will help us proceed from research to clinical applications of optical modalities for research, diagnosis, and treatment of stroke.

1. Introduction

A stroke occurs when partial blood supply to brain suddenly stops or when the blood vessels of the brain burst, consequently spilling blood into a space surrounding the brain cells which die when they no longer receive oxygen and nutrients from the blood or when there is sudden bleeding in or around the brain. Stroke has two forms: (1) ischemic blockage in blood vessels of the brain, and (2) hemorrhagic stroke that causes bleeding in the brain. Globally, stroke is the second-largest cause of death and third-largest cause of disability [1]. In 2019, a number of global stroke patients reached 115 million. A number of ischemic stroke patients was 77.2 million, a number of hemorrhagic stroke patients was 20.7 million, and a number of stroke patients due to subarachnoid hemorrhage was 8.4 million [2]. Also, 3.3 million people worldwide died from stroke in 2019.
To look a little further at the progression of stroke in the cell, tissue, and brain function, first, clogging or bursting blood vessels results in irreversible damage to the center of the brain tissue directly associated with it. After that, brain damage and penumbra occur in the surrounding brain tissues. In neuronal cells of tissues damaged by stroke, glutamate is released from ion channel collapse and excessive inflow of calcium occurs [3,4]. In addition, energy production decreases due to a damage of DNA (deoxyribonucleic acid) [5,6]. As a result, both necrosis and apoptosis of neuronal cells occur in brain tissues damaged by stroke. Also, blood corpuscles are accumulated by factors such as interleukin 8 (IL-8) expressed in damaged brain tissue by stroke [7]. It enhances an inflammatory response of the brain tissue. In such damaged brain tissues by stroke, neurons and blood vessels are regenerated and recovery proceeds for weeks to months [8]. Studies to understand anatomical and functional brain damages related to complex connectivity and study functional recovery mechanisms due to neuroplasticity after stroke have been actively conducted [9,10,11]. Research, development, and medical applications in the diagnosis, prevention, treatment immediately after stroke, and rehabilitation/therapeutic techniques for post-stroke patients have extensively progressed for high incidence and mortality of stroke and studies on mechanism of stroke, recovery processes, and therapeutic techniques developments.
Biomedical optics is a research field that studies both the principle of light-biological specimen interactions and development of medical applications [12,13]. Research and diagnosis techniques using light have advantages such as lower invasiveness and the ability to obtain varied biological and medical information compared to diagnostic modalities using radiation and other energies. Additionally, benefits of light-based therapeutic methods have fewer side effects and can be combined with conventional treatments such as chemical drugs. In particular, neurophotonics has been extensively used in studies using light for research, diagnostic, and therapeutic modalities in brain and nervous system diseases [14,15].
In this article, we explored optical techniques for research, diagnostics, and therapeutic modalities for stroke and related brain injuries. Specifically, three detailed techniques are addressed: (1) photochemical thrombosis to establish stroke animal models with localized brain lesions; (2) optical imaging techniques for in vivo preclinical studies for understanding stroke mechanism and developing diagnostic or treatment modalities; and (3) optical neurostimulation based therapy of stroke and preclinical and clinical assays related to it; and. An exploration, an identification, and an analysis of previous studies of optical modalities for stroke are confident that they will help with the processes from studies to clinical applications of optical methods for research, diagnosis, and treatment of stroke.

2. Investigating In Vivo Stroke Model Using Photochemical Thrombosis

2.1. Photochemical Thrombosis-Based Stroke Animal Model Establishments

Each of research, diagnosis and treatment modalities require verification, especially preclinical and clinical tests, which are mandatory for diagnosis and treatment devices that are to be employed for medical purposes. Development of a stroke-induced animal model for preclinical tests and modalities to build such an animal model are being extensively investigated [16,17,18,19]. A typical stroke animal model investigation is achieved either through surgical blocking of specific vessels [20,21,22] or through a biochemical method [23,24]. However, stroke in local brain regions is investigated using photochemical thrombosis. An investigation of a photochemical stroke animal model is based on a generation of blood clots using rose bengal [25,26,27,28]. Rose bengal reacts with a green light irradiation to produce reactive oxygen species, which then activate factors that are involved in a coagulation of blood and production of thrombosis. To use this characteristic of rose bengal, an animal’s blood vessel is first injected with rose bengal, and after waiting for it to spread to all of the blood vessels, a green light is irradiated in a desired brain region, which causes photochemical thrombosis and stroke in the desired area. This section describes several studies focused on photochemical thrombosis investigation modalities for performance enhancements such as an improvement of location accuracy of brain damages and an inhibition of excessive brain damage.
A. Sigler et al. investigated a photothrombosis inducing module that was integrated with an upright microscope to generate a photothrombotic stroke on a single brain arteriole [29]. To be specific, a 532-nm laser diode module and an optical cage system were applied for a construction of the photothrombosis inducing module, which was connected to a filter slot of an upright microscope. Moreover, this module provides a function that can form blood clots in certain cerebral blood vessels while observing them under a microscope. When laser irradiation on a single brain vessel in mice was maintained for 120 s, a simultaneous blockage of a targeted brain vessel was verified using microscopic images (Figure 1a). In addition, laser speckle imaging was utilized to confirm the vascular obstruction caused by photothrombotic stroke. For a similar purpose, M. Alaverdashvili et al. developed a reconfigured laser system to reduce a deviation in induced sizes of brain damages caused by photothrombosis [30]. In the laser system that they proposed, a polarizer or a variable neutral density filter were employed to maintain a constant output irradiation power of laser, which was verified using an optical power meter. An in vivo study on photothrombosis inducing in rats’ brains and 2,3,5-Triphenyltetrazolium chloride (TTC) staining used for estimating locations and depths of photothrombotic strokes confirmed that the system with a maintained optical irradiation power is useful for inducing a stroke and the consequent brain damage of a certain size. U.I. Tuor et al. investigated a minor photothrombotic stroke animal model using a fiber-optic laser irradiation probe to generate Rose Bengal photoactivation [31]. In detail, a probe with a halogen lamp consisting of an optical fiber with a 13-mm diameter, a heat filter, and collimator was constructed to induce a photothrombotic stroke. According to magnetic resonance images described in Figure 1b, an investigation of a small animal model with minor stroke was achieved using a fiber-optic photothrombosis inducing probe. A photothrombosis inducing device integrated with a stereotaxic frame was also developed to produce brain damage at a certain position [32,33]. The stereotaxic frame, which was applied in this study, was magnetic resonance compatible (i.e., it was configured to move with an animal’s location to acquire magnetic resonance images of the brain). Additionally, it provided a function to link indicators of the stereotaxic frame and coil to provide support for image acquisitions, so that it could be held firmly, ensuring the least possible movement.
Various optical imaging modalities were utilized to analyze stroke and its resultant brain damages in cell tissue. The preclinical animal studies were discussed in a previous section. Relevant studies have also been conducted extensively in which these optical imaging techniques are applied to methods used for verifying stroke-inducing animal models or for analyzing changes in lesions or treatments. Q. Liu et al. used a real-time and high-resolution laser speckle brain imaging system to confirm a blockage of blood vessels and vascular changes in the brain of a rodent model using photochemical thrombosis [34]. In this study, changes in a specific vessel over time were quantitatively explored through high-speed imaging of brain vessels. S. Zhang and T.H. Murphy employed in vivo two-photon fluorescence microscopy to observe changes in dendrites as well as a distribution of blood vessels in a photothrombotic animal (mice) model [35]. In addition, it was confirmed that changes in neurons and biomolecules could also be observed using in vivo two-photon imaging, and precise measurements were achieved for damage and changes in dendrites through photochemical thrombosis, as illustrated in Figure 2. H. Lu et al. investigated a miniaturized and lightened infrared (780 nm) vessel imaging device, which was equipped with a combination of fiber-coupled blue and green lasers. The device is designed to generate photochemical thrombosis and acquire a brain vessel image of a freely moving animal [36]. The miniaturized imaging device is significant in that it can immediately monitor effects of photochemical thrombosis and changes in cerebral blood vessels without a need for location-limited noninvasive brain imaging techniques such as an MRI.

2.2. Summary of In Vivo Stroke Model Investigations

In this section, we explored several studies and techniques to establish photothrombosis animal models. Compared to surgical stroke animal model establishment methods, the photothrombosis has advantages of being relatively easy to construct stroke animal models, less likely to die from important vascular obstruction, and easier to control a location of localized stroke. When an establishment of standardized experimental devices and protocols is further carried out by follow-up studies, it is expected that the photothrombosis can be applied to secure a certain stroke animal model and use it for research of various diagnostic and treatment techniques.

3. Optical Imaging Modalities to Develop and Verify Diagnosis/Treatment Techniques of Stroke

3.1. Fluorescence Imaging

Fluorescence imaging is a modality that measures biological, chemical, or biophysical information and its time-varying changes by imaging fluorescence indicators that emit fluorescent signals including required information. Moreover, it has an advantage of non-invasive acquisition of functional and specific bio-information without destroying any biomaterials, for instance, cells [37,38,39,40,41]. For this reason, fluorescence imaging is utilized in both in vitro and in vivo preclinical assays for understanding their causes and verifying diagnostic/therapeutic techniques of stroke by imaging the fluorescent markers that display biochemical and biophysical information related to strokes and its consequent brain damages. In particular, structural neuroplasticity [42,43], axonal regrowth and sprouting [44,45], a regeneration of dendrites [46] in recovery and rehabilitation after stroke were understand by high-resolution fluorescence imaging techniques. In this section, several studies on in vivo animal studies on stroke using fluorescence imaging were explored.
Confocal and two-photon fluorescence microscopy are some of the advanced fluorescence imaging techniques that overcome limits of penetration depth in general fluorescence imaging with a higher resolution [47,48,49,50]. Therefore, these have been employed to analyze stroke-induced in vivo animal models and conduct preclinical tests of stroke-related medical devices. N. Nishimura et al. identified blood clots, which were generated by photothrombosis, in arteries penetrating the longitudinal axis, and changes in blood clots and surrounding blood vessels by observing fluorescein/dextran labeled plasmas using two-photon microscopy [51]. L.T. Watts et al. obtained fluorescence images of blood clots formed inside a cerebral blood vessel beyond a thinned skull of mice [52]. In this study, thrombosis in cerebral blood vessels was confirmed by obtaining fluorescence images of injected fluorescein/dextran in the vessels using a confocal fluorescence microscope. Brain damages from a generation of thrombosis were also verified by 2,3,5-triphenyltetrazolium chloride (TTC) staining after an extraction of the brain. C.J. Schrandt et al. conducted a study to analyze a transient progression of a distribution of the cerebral blood vessels and blood flows using optical speckle and two-photon fluorescence microscopes [53]. The multi-exposure optical speckle imaging used in this study showed a possibility of a long-term analysis of cerebral blood vessels in stroke-induced animal models through its comparison with the two-photon fluorescence imaging.
Studies that analyze strokes and a resultant brain damage using fluorescence imaging of in vivo brain of animal models mainly employed near-infrared probes, which correspond to one of the highest transmission windows [54]. G. Hong et al. developed an optical imaging system and technique to acquire a near-infrared image of a distribution of blood vessels in a mouse brain without removing a skull using a carbon nanotube indicator that emits functionalized optical signals between 1300 and 1400 nm [55]. This imaging technique can obtain arterial and venous veins in the mouse brain with minimal invasiveness up to a depth of approximately 2 mm. Moreover, an experiment to observe vascular loss in a local brain region, which is caused by middle cerebral artery occlusion (MCAO) induced stroke, confirmed its applicability as a stroke research tool. Y. Zhang et al. applied a near-infrared fluorescence indicator to noninvasive acquisitions of fibrin protein deposition in stroke-induced mice as illustrated in Figure 3a [56]. A comparative study of damaged brain areas for ex vivo brain slices confirmed that the fluorescence indicator and noninvasive near-infrared imaging based on it has sufficient correlation to indicate lesions caused by a stroke.

3.2. Functional Near-Infrared Spectroscopy

Both ischemic and hemorrhagic strokes are diseases caused by changes and abnormalities in a distribution of blood in cerebral blood vessels. Therefore, methods to extract information on cerebral blood flows and hemoglobin distributions in cerebral blood vessels, for instance, Doppler ultrasound imaging and vascular magnetic resonance imaging (MRI), can be employed as a noninvasive diagnostic method for stroke. The optical imaging technology, defined as a functional near-infrared spectroscopy (fNIRS), provides an ability to obtain distribution information of oxy-/deoxy-hemoglobin and total blood concentrations using infrared light sources [57,58,59,60]. fNIRS does not use specific indicators, unlike fluorescence imaging. Therefore, it has been applied to preclinical tests as well as to brain function analysis and a verification of stroke rehabilitation at a clinical stage.
Y. Shang et al. investigated a diffuse optical oxy-/total hemoglobin monitoring system using infrared light sources with two peak wavelengths of 785 nm and 854 nm [61]. The system, which consists of two light sources and detectors, can measure hemoglobin concentrations and blood flows in left and right sides of a mouse brain. Moreover, it demonstrated high correlation in a comparative study using a laser Doppler flowmeter and a mouse model with repeated ischemia. C.H. Han et al. applied fNIRS to confirm cerebral hemodynamic responses derived using transcranial electrical stimulation, which is one of the neurostimulation-based therapeutic techniques for brain diseases including stroke [62]. Two light sources with central wavelengths of 690 nm and 830 nm, and detectors associated with fiber-optic probes, were installed in left and right sides of a rat brain. Further, hemodynamic responses were measured for direct current simulation by an electrode, which was placed between the fiber-optic probes of the light sources and the detector. Z.J. Lin et al. applied an infrared diffusion tomography to establish three-dimensional hemodynamic changes to a rat model that performed ischemic stroke (derived by middle cerebral artery to create a permanent occlusion (MCAO)) and its treatment (reperfusion), as illustrated in Figure 3b [63].
fNIRS is also applied to a diagnosis of stroke and to monitoring an effectiveness of rehabilitation and treatment in cerebrovascular stroke patients in a clinical stage. M.J.H. Aries et al. conducted a pilot study of fNIRS to detect acute ischemic stroke and a good responsiveness in stroke-derived changes was confirmed in a small clinical research study [64]. G. Giacalone et al. studied a diagnostic feature extraction of acute ischemic stroke for oxy-/deoxy-/total hemoglobin and tissue oxygen saturation that can be measured by fNIRS (with detailed probe distribution described in Figure 3c) in a clinical stage [65]. Several research groups analyzed a validity and performance of rehabilitation applications and gait training for stroke through a noninvasive observation of cortical activations using fNIRS [66,67,68,69,70]. P.-Y. Lin et al. studied typical cortical activations during a cycle rehabilitation exercise of stroke patients [71]. The study found that despite the existing differences in a degree of activation, both passive and active cycle rehabilitation exercises contribute to premotor cortex activations. Studies using fNIRS as a clinical validation tool for stroke treatments based on various physical stimulations are also being conducted extensively. A. Dutta et al. measured electroencephalographic signals and oxy-hemoglobin during transcranial direct current electrostimulation therapies in stroke patients using a multimodal neuro-signal measurements probe linked to electroencephalography (EEG) and fNIRS [72,73]. From the measurements, variations in cerebrovascular coupling due to the electrical brain stimulation were observed and a potential for development of devices capable of brain stimulation and simultaneous monitoring of multimodal neural signals for treatment efficiency verification was identified. G. Litscher et al. conducted a study using fNIRS to explore brain activities and its responses when a laser stimulation and traditional acupuncture treatments were applied in parallel to stroke patients [74]. Although precisions and evidences should be reinforced through follow-up clinical studies, variations in the brain activity in local areas were noninvasively observed using the multimodal stimulations by fNIRS.
fNIRS was commercialized as a cerebral hematoma diagnostic device at an emergency site due to an advantage that the location of a hematoma can be searched quickly and noninvasively [75,76]. Typically, a clinical study was conducted to diagnose bleeding and the hematoma in the brain, using a commercialized fNIRS-based intracranial hematoma diagnostic device (Infrascanner 2000, InfraScan, Inc., Philadelphia, PA, USA) at helicopter emergency medical services in The Netherlands [77].
Figure 3. (a) Near-infrared images of in vivo fluorescence using fibrin deposition in stroke-induced mice at 6, 8, 20, and 96 h after stroke was triggered. Reprinted with permission from Ref. [56]. Copyright 2012 Public Library of Science. (b) Near-infrared optical diffusion tomographic rat brain images of oxygenate/deoxygenate/total hemoglobin concentrations 30 min after an ischemic stroke. Reprinted with permission from Ref. [63]. Copyright 2014 Elsevier B.V. (c) fNIRS probe arrays designed for an analysis of acute ischemic stroke patients in clinical tests and computed tomographic images used to determine a measurement possibility for areas with and without brain lesions. Reprinted with permission from Ref. [65]. Copyright 2019 International Society for Optics and Photonics (SPIE).
Figure 3. (a) Near-infrared images of in vivo fluorescence using fibrin deposition in stroke-induced mice at 6, 8, 20, and 96 h after stroke was triggered. Reprinted with permission from Ref. [56]. Copyright 2012 Public Library of Science. (b) Near-infrared optical diffusion tomographic rat brain images of oxygenate/deoxygenate/total hemoglobin concentrations 30 min after an ischemic stroke. Reprinted with permission from Ref. [63]. Copyright 2014 Elsevier B.V. (c) fNIRS probe arrays designed for an analysis of acute ischemic stroke patients in clinical tests and computed tomographic images used to determine a measurement possibility for areas with and without brain lesions. Reprinted with permission from Ref. [65]. Copyright 2019 International Society for Optics and Photonics (SPIE).
Applsci 12 01891 g003

3.3. Optical Coherence Tomography

Optical coherence tomography (OCT) is an imaging technique that employs an infrared interferometer to obtain optical interference information in depth using light scattering materials and scans it to acquire two- or three-dimensional images of optical coherence information [78,79,80]. An imaging resolution of OCT is typically in micron units, and OCT is generally used as a tool and a diagnostic device for obtaining biomedical information such as eyes (retina) and skin [81,82,83,84]. A more advanced form of OCT, which can acquire bio-information in deeper areas through light sources with a longer wavelength and can be used for acquiring functional information such as vascular distributions through high-speed signal measurements and post-processing techniques, has been developed, which increases a scope of its applications [85,86,87,88]. In this section, a series of studies that conducted studies of cerebrovascular stroke by analyzing information about blood vessels and blood flow rates using advanced OCT instruments are discussed.
L. Yu et al. investigated a Doppler OCT to measure blood circulation in ischemic stroke-induced mice [89]. The study employed a 532 nm LASER to induce an ischemic stroke and changes in blood flow in a localized area, where stroke was generated, were observed using spectral Doppler OCT. E. Osiac et al. analyzed structural brain damages induced by artificial stroke [90]. The rat brain was extracted after different time periods after stroke was induced, and areas of brain damages were imaged using OCT. The results suggested a possibility of OCT as a pathological analysis equipment for ex vivo brain tissues with an acute or chronic stroke. As a follow-up study, an application of an image analysis based on artificial neural networks was additionally applied to classifications of stroke injuries in ex vivo brain tissues [91]. V.J. Srinivasan et al. employed a depth-resolved OCT angiographic imaging, which was integrated in a conventional upright microscope, to analyze acute cerebral ischemic stroke, which was induced by distal MCAO, and a chronic restoration of blood vessels over time in in vivo small animal studies, as illustrated in Figure 4 [92]. S. Chen et al. applied depth-coded visible OCT angiography for making accurate analyses of microvascular blood flow variations when a localized ischemic stroke was induced through photothrombosis [93,94]. In this study, the difference in the blood oxygen saturation caused by a localized ischemic stroke induction was also observed simultaneously by calculating oxygen saturations using the spectroscopic images from visible OCT angiography.

3.4. Photoacoustic Imaging and Tomography

Photoacoustic imaging is a label-free bio-imaging technique in which minute thermal expansion and its consequent pressure changes, which are caused by excitations of pulsed light to a specific biomaterial, are measured in a range of acoustic or ultrasonic waves, and are scanned in multiple dimensions to obtain information on a distribution of a particular biomaterial [95,96]. In detail, photoacoustic imaging consists largely of five steps to obtain images: pulsed light irradiation to excite a specific material, excitation → relaxation, increasing a pressure, measurement of the increased pressure, and finally, image reconstruction [97]. This technique is widely employed in in vivo studies due to its advantages such as deeper imaging depth compared to that of optical imaging, which gives and receives optical signals, and it is label-free [98,99,100,101]. It also exhibits potential for application in minimally invasive clinical diagnosis of several diseases such as breast cancer and tumors [102,103]. As a research tool for stroke, several studies were conducted to analyze photoacoustic signals generated by oxy-/deoxy-hemoglobin for understanding processes of stroke and verifying diagnostic/therapeutic techniques.
J. Laufer et al. developed a high-resolution photoacoustic imaging system that allows precise, three-dimensional observation of blood vessels in mice brains [104]. The system consists of a nanosecond pulsed LASER with excitation wavelength ranges between 590 nm and 889 nm, and an interferometric detection film as a photoacoustic signal sensor. Moreover, it provides images of brain blood vessels with a high sensitivity, precision and greater image acquisition depth. L. Lin et al. investigated a high-speed photoacoustic imaging system to acquire time-varying distributions of blood vessels in mice brains [105]. The photoacoustic microscopy measured a loss of vessels caused by micro-hematoma and laser irradiation using precise photoacoustic brain blood vessels image acquisition, and an analysis of shrinkage response of cerebral blood vessels, which was carried out by injection of phenylephrine, was also performed using photoacoustic imaging. M. Kneipp et al. employed a multispectral optoacoustic imaging to acquire oxy-/deoxy-hemoglobin distributions of mice brains when an artificial stroke was induced by a MCAO, as illustrated in Figure 5 [106]. The imaging method using this proposed optoacoustic imaging system is meaningful in that it provides functional hemodynamic images with enhanced depth ranges while ensuring noninvasiveness. Z. Deng et al. monitored responses of blood vessel distributions and blood oxygen saturations in a rat brain after an induction of acute stroke using a multimodal imaging system, which consisted of laser speckle and photoacoustic imaging [107]. D. Wu et al. developed a multispectral photoacoustic imaging system for a noninvasive analysis of intracerebral hematoma locations induced by collagenase [108]. Images computed using photoacoustic signals at 532 nm, 750 nm, and 875 nm as excitation wavelengths offered location and size information of a lesion in hematoma and perihematomal regions. R. Ni et al. applied multispectral photoacoustic imaging to acquire noninvasive measurements of a matrix-metalloproteinase, a unique group of enzymes associated with acute stroke and brain damages, in in vivo mouse model of an ischemic stroke [109].

3.5. Summary of Optical Imaging Modalities

In this section, we looked at four optical imaging techniques (fluorescence imaging, fNIRS, OCT, and photoacoustic imaging) to study and diagnosis stroke. Optical imaging employed in clinical applications is fNIRS. Fluorescence imaging could be limited in use in clinical sites due to limitations in clinical compatibility of fluorescence staining markers, and it is expected to be actively used in cellular and preclinical studies of stroke. Compared with ultrasound and magnetic resonance imaging, OCT and photoacoustic imaging clearly have limitations in penetrations depth, but are expected to be used in clinical applications through further researches to improve the performance. In addition, OCT and photoacoustic imaging can be employed in combination with fluorescence imaging for cellular and preclinical studies. An application of image post-processing and artificial intelligence-based data analysis techniques will help each optical imaging technique provide more meaningful information for research and diagnosis of stroke.

4. Light Therapy for Mitigation/Treatment of Stroke

4.1. Light Therapy for Stroke Based on Photobiomodulation

Photobiomodulation means a use of light in selected wavelengths for purposes of healing tissues, enhancing a recovery speed, alleviating pain, or restoring function [110,111,112,113]. Among several types of photobiomodulation, a method of irradiating light to the brain for performing functions of treatments, recovery and symptom mitigation is called transcranial photobiomodulation [114,115,116,117,118], and light with spectral bands of red or infrared is generally employed in this process due to its relatively high penetration efficiency. Several research groups have studied transcranial photobiomodulation for brain diseases in three categories: stroke/ischemia, degenerative diseases (such as Alzheimer’s disease), and psychiatric diseases (for instance, bipolar disorder). This section explores the latest progress and important points regarding the studies focused on transcranial photobiomodulation to stroke and its consequent brain injuries.
Several research groups conducted preclinical studies on small and middle-sized animals to identify effects of therapeutic optical stimulation on treatment of stroke and symptoms related to it. L. DeTaboada et al. explored a possibility of recovering neurological deficits caused by a stroke using low-level laser therapy with a peak wavelength of 808 nm in a rat stroke model, which was established using a middle cerebral artery to create a permanent occlusion (MCAO) [119]. The quantitative estimation of neurological deficits from six behavioral tests confirmed that the therapeutic photostimulation on both sides (i.e., a stroke-induced side and on an opposite side improved neurological deficits after stroke). An immunohistochemical analysis of neurons using Bromodeoxyuridine and Tubulin isotype III also confirmed a decrease in neurological damages using therapeutic photostimulation with an 808 nm laser at 7.5 mW/cm2 [120]. Preclinical tests using a rabbit stroke model, where stroke was caused by artificially injecting blood clot particles, also indicated that pulsed near-infrared laser irradiations aid healing of brain damages [121]. In this study, a significant impact was observed on a treatment of pulsed photostimulation under two different conditions (1000 Hz with a duty cycle of 30% and 100 Hz with a duty cycle of 20%, with a power density = 7.5 mW/cm2) after 6 h of stroke. However, no therapeutic effect of continuous laser irradiation could be observed when compared statistically with control results, as illustrated in Figure 6. H.I. Lee et al. employed a light emitting diode (LED), with 610 nm as a peak wavelength, in a preclinical study of photostimulation-based stroke therapies [122]. Furthermore, light irradiation using the LED-integrated medical device prototype was applied to a photothrombotic mouse stroke model, and a possibility of using the LED-integrated medical device to treat stroke and mitigate resultant symptoms was confirmed through a comparison of a volume of ischemic brain damages with a control model that was untreated. Moreover, a functional recovery from treatments through the LED-integrated medical device was verified using immunofluorescence assays (Bromodeoxyuridine and an astrocyte marker) [123]. B. Argibary et al. examined preclinical tests to apply infrared (830 nm) LEDs to treatments based on photobiomodulation after an ischemic stroke [124]. Specifically, these researchers applied the therapeutic photostimulation to rats that were subjected to stroke by a transient MCAO, and employed functional MRI (fMRI) to explore functional recoveries of the brain area where ischemic stroke was induced. Although the study could not identify a recovery of stroke-induced lesions and functional reduction due to stroke by LED-based optical neurostimulation, it confirmed that fMRI is a sufficient modality to validate an efficiency of therapeutic photostimulation, both structurally and functionally. Additionally, it was suggested that an application of efficient light irradiation protocols and therapeutic photostimulation for stroke damages in smaller areas should be further studied using fMRI.
Clinical studies for stroke patients and a possibility of noninvasive treatments based on photobiomodulation have been progressed for a treatment of stroke. Y. Lampl et al. reported that therapeutic infrared photobiomodulations had successful outcomes in clinical trial studies with 120 ischemic stroke patients [125,126]. This clinical trial dealt with infrared laser treatments immediately after stroke or within 24 h from the occurrence of stroke, and researchers confirmed that treatments conducted using photo-therapy within 24 h had a satisfactory effect. Moreover, clinical trial studies and validity analyses of transcranial photostimulation to treat an acute ischemic stroke were later conducted on a larger scale, covering 780 patients in 5 countries [127]. H.L. Casalechi et al. conducted clinical studies on an effectiveness of a therapy combining photobiomodulation with magnetic stimulations to a muscle and a nerve for restoring functional mobilities of stroke patients [128]. It has been confirmed that the magnetic stimulation (30 J per site) with light therapy using four infrared diodes (905 nm), pulsed laser diodes, four infrared (with 875 nm as their peak wavelength) LEDs, and four red (640 nm as a peak wavelength) LEDs can significantly recover the functional mobility of stroke patients.

4.2. Optogenetic Neurostimulation to Treat Stroke

Meanwhile, several researchers have developed a technique to treat stroke using optogenetic-based neurostimulation at a preclinical stage and secured a basis for stroke treatment by optogenetic brain stimulations. Optogenetics is a biological technique that genetically ion channels of neurons so that they can be opened or closed in response to light [129,130,131]. So, the neurons modified by optogenetic techniques can be activated or inhibited by light. M.Y. Cheng et al. suggested an optogenetic stimulation as a method to promote functional recovery [132]. The result of this study indicated that when channelrhodopsin-2 was expressed to neurons in an ipsilesional brain region of mice that caused stroke and optical stimulation was induced, blood flow and a neurovascular response increased. M. Song et al. introduced that optogenetic stimulations promoted a neurogenesis of stroke mice [133]. A.M. Shah et al. conducted a preclinical study that optical stimulations of neurons in lateral cerebellar nucleus can contribute to brain function recovery after stroke [134]. A follow-up study by the same research group confirmed that optogenetic neurostimulations reduced neuronal nitric oxide synthase, one of the regulators of neurovascular responses in stroke [135]. It was confirmed that this decrease in neuronal nitric oxide synthase causes recovery of functional degradation by stroke. A preclinical in vivo study was also conducted that both optogenetic neurostimulations and rehabilitation training. E. Conti and co-researchers confirmed that brain function degradation caused by stroke was be more recovered when optogenetic neuronal activations to a motor cortex and foreleg exercise training were applied together to mice with an induction of stroke and an expression of channelrhodopsin-2 [136].

4.3. Summary of Optical Treatments for Stroke

In this section, we explored two types of optical therapeutic techniques (photobiomodulation-based treatment and optogenetic neurostimulation) for stroke. Photobiomodulation-based stroke treatment is highly likely to be employed as a non-drug therapeutic tool for brain diseases, and can be applied in parallel with conventional drugs. Based on the merits of non-causing unexpected problems by spreading through a body such as conventional drugs and low-cost, photobiomodulation-based treatments for brain diseases are expected to be applied in clinical sites and commercialized step by step. On the other hand, additional developments are needed to understand mechanisms and transmit light beyond a skull efficiently. Optogenetic neurostimulations to treat stroke were actively studied at a preclinical in vivo stage for a purpose of understanding a treatment mechanism and verifying effectiveness. The optogenetic stroke treatment showed potential for application as a therapeutic technique for stroke in parallel with conventional drugs. A limitation to be solved for clinical applications is that the safety and stability of optogenetic probes to express neurons must be secured.

5. Conclusions

In this article, optical modalities for research, diagnosis, and treatment of stroke and the consequent brain damage were described. The optical imaging techniques employed for studying cerebrovascular strokes, such as fluorescence imaging, fNIRS imaging, which is based on light irradiations with multiple near-infrared wavelengths for noninvasive acquisition of hemoglobin information; OCT imaging, which is used to obtain information of label-free brain tissue and blood vessels; and photoacoustic imaging employed for studying cerebral blood vessels and to gather functionalized information were explored. Each of these techniques is extensively employed for research and in diagnostic areas. In particular, fNIRS has been employed as a medical device for a diagnosis of acute cerebral hemorrhage and consequent brain damage at emergency sites. Stroke treatments through optical neurostimulation are extensively researched in preclinical studies using animal models and comparative exploration clinical studies. When the results and prospects of these studies are identified, we expect that practical applications and commercialization would be phased out through technological advancements and large-scale clinical verifications. The establishment of stroke animal model, which is based on photothrombosis, is well understood. Moreover, uniform stroke inducement in terms of optical system investigation and upgrade have been identified in previous studies. We are sure that quantification and verification modalities during the establishment of a stroke animal model and stroke-derived brain damages are essential for preclinical studies focused on diagnostic and therapeutic techniques of stroke. Moreover, these optical techniques could be extensively utilized in developments of diagnostic and therapeutic modalities for other brain diseases such as Alzheimer’s disease and certain psychiatric disorders.

Author Contributions

Conceptualization, S.S.O. and J.-r.C.; writing—original draft preparation, S.S.O., Y.K. and J.-r.C.; writing—review and editing, S.S.O., Y.B.L., S.K.B., J.S.K. and S.-h.A.; visualization, S.S.O., Y.K. and J.-r.C.; supervision, S.S.O. and J.-r.C.; project administration, J.-r.C.; funding acquisition, S.S.O. and J.-r.C. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported from a grant (NRF-2020R1C1C1012230) of the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science & ICT. This work was also supported by a grant (HI17C1501) of the Health Technology R&D Project through the Korea Health Industry Development Institute (KHIDI) funded by the Ministry of Health & Welfare, Republic of Korea.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Johnson, W.; Onuma, O.; Owolabi, M.; Sachdev, S. Stroke: A global response is needed. Bull. World Health Organ. 2016, 94, 634. [Google Scholar] [CrossRef] [PubMed]
  2. 2021 Heart Disease & Stroke Statistical Update Fact Sheet—Global Burden of Disease, American Heart Association. 2021. Available online: https://www.heart.org/-/media/PHD-Files-2/Science-News/2/2021-Heart-and-Stroke-Stat-Update/2021_Stat_Update_factsheet_Global_Burden_of_Disease.pdf (accessed on 2 February 2022).
  3. Sims, N.R.; Muyderman, H. Mitochondria, oxidative metabolism and cell death in stroke. Biochim. Biophys. Acta 2010, 1802, 80–91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Mattson, M.P.; Culmsee, C.; Yu, Z.F. Apoptotic and antiapoptotic mechanisms in stroke. Cell Tissue Res. 2000, 301, 173–187. [Google Scholar] [CrossRef] [PubMed]
  5. Li, X.; Klaus, J.A.; Zhang, J.; Xu, Z.; Kibler, K.K.; Andrabi, S.A.; Rao, K.; Yang, Z.-J.; Dawson, T.M.; Dawson, V.L.; et al. Contributions of poly(ADP-ribose) polymerase-1 and -2 to nuclear translocation of apoptosis- inducing factor and injury from focal cerebral ischemia. J. Neurochem. 2010, 113, 1012–1022. [Google Scholar] [CrossRef] [Green Version]
  6. Abdelkarim, G.E.; Gertz, K.; Harms, C.; Katchanov, J.; Dirnagl, U.; Szabo, C.; Endres, M. Protective effects of PJ34, a novel, potent inhibitor of poly(ADP-ribose) polymerase (PARP) in in vitro and in vivo models of stroke. Int. J. Mol. Med. 2001, 7, 255–260. [Google Scholar] [CrossRef]
  7. Domac, F.M.; Misirli, H. The role of neutrophils and interleukin-8 in acute ischemic stroke. Neuroscience 2008, 13, 136–141. [Google Scholar]
  8. Arvidsson, A.; Collin, T.; Kirik, D.; Kokaia, Z.; Lindvall, O. Neuronal replacement from endogenous precursors in the adult brain after stroke. Nat. Med. 2002, 8, 963–970. [Google Scholar] [CrossRef] [Green Version]
  9. Pin-Barre, C.; Hugues, N.; Constans, A.; Berton, E.; Pellegrino, C.; Laurin, J. Effects of different high-intensity interval training regimens on endurance and neuroplasticity after cerebral ischemia. Stroke 2021, 52, 1109–1114. [Google Scholar] [CrossRef]
  10. Cui, J.; Kim, C.S.; Kim, Y.; Sohn, M.K.; Jee, S. Effects of repetitive transcranial magnetic stimulation (rTMS) combined with aerobic exercise on the recovery of motor function in ischemic stroke rat model. Brain Sci. 2020, 10, 186. [Google Scholar] [CrossRef] [Green Version]
  11. Dihne, M.; Grommes, C.; Lutzenburg, M.; Witte, O.W.; Block, F. Different mechanisms of secondary neuronal damage in thalamic nuclei after focal cerebral ischemia in rats. Stroke 2002, 33, 3006–3011. [Google Scholar] [CrossRef] [Green Version]
  12. French, P. Biomedical optics. Phys. World 1999, 12, 41. [Google Scholar] [CrossRef]
  13. Wang, L.V.; Wu, H.I. Biomedical Optics: Principle and Imaging; John Wiley & Sons: Hoboken, NJ, USA, 2012. [Google Scholar]
  14. Torricelli, A.; Contini, D.; Mora, A.D.; Pifferi, A.; Re, R.; Zucchelli, L.; Caffini, M.; Farina, A.; Spinelli, L. Neurophotonics: Non-invasive optical techniques for monitoring brain functions. Funct. Neurol. 2014, 29, 223–230. [Google Scholar] [CrossRef] [PubMed]
  15. Choi, Y.K.; Zheng, G.; Augustine, G.J.; Hochbaum, D.; Cohen, A.; Knöpfel, T.; Pisanello, F.; Pavone, F.S.; Vellekoop, I.M.; Booth, M.J.; et al. Roadmap on neurophotonics. J. Opt. 2016, 18, 093007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Dittmar, M.; Spruss, T.; Schuierer, G.; Horn, M. External carotid artery territory ischemia impairs outcome in the endovascular filament model of middle cerebral artery occlusion in rats. Stroke 2003, 34, 2252–2257. [Google Scholar] [CrossRef] [PubMed]
  17. Ansari, S.; Azari, H.; McConnell, D.J.; Afzal, A.; Mocco, J. Intraluminal middle cerebral artery occlusion (MCAO) model for ischemic stroke with laser doppler flowmetry guidance in mice. J. Vis. Exp. 2011, 51, e2879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Fluri, F.; Schuhmann, M.K.; Kleinschnitz, C. Animal models of ischemic stroke and their application in clinical research. Drug Des. Dev. Ther. 2015, 9, 3445–3454. [Google Scholar]
  19. Selvamani, A.; Sathyan, P.; Miranda, R.C.; Sohrabji, F. An antagomir to microRNA Let7f promotes neuroprotection in an ischemic stroke model. PLoS ONE 2012, 7, e32662. [Google Scholar] [CrossRef] [Green Version]
  20. Yanamoto, H.; Nagata, I.; Niitsu, Y.; Xue, J.H.; Zhang, Z.; Kikuchi, H. Evaluation of MCAO stroke models in normotensive rats: Standardized neocortical infarction by the 3VO technique. Exp. Neurol. 2003, 182, 261–274. [Google Scholar] [CrossRef]
  21. Bederson, J.B.; Pitts, L.H.; Tsuji, M.; Nishimura, M.C.; Davis, R.L.; Bartkowski, H. Rat middle cerebral artery occlusion: Evaluation of the model and development of a neurologic examination. Stroke 1986, 17, 472–476. [Google Scholar] [CrossRef] [Green Version]
  22. Fan, J.; Li, Y.; Fu, X.; Li, L.; Hao, X.; Li, S. Nonhuman primate models of focal cerebral ischemia. Neural Regen. Res. 2017, 12, 321–328. [Google Scholar]
  23. Abeysinghe, H.C.S.; Roulston, C.L.; Abeysinghe, H.C.S.; Roulston, C.L. A complete guide to using the Endothelin-1 model of stroke in conscious rats for acute and long-term recovery studies. In Traumatic and Ischemic Injury, 1st ed.; Tharakan, B., Ed.; Humana Press: New York, NY, USA, 2018; pp. 115–133. [Google Scholar]
  24. Chan, H.H.; Cooperrider, J.L.; Park, H.J.; Wathen, C.A.; Gale, J.T.; Baker, K.B.; Machado, A.G. Crossed cerebellar atrophy of the lateral cerebellar nucleus in an endothelin-1-induced, rodent model of ischemic stroke. Front. Aging Neurosci. 2017, 9, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Rosenblum, W.I.; El-Sabban, F. Platelet aggregation in the cerebral microcirculation: Effect of aspirin and other agents. Circ. Res. 1977, 40, 320–328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Watson, B.D.; Dietrich, W.D.; Busto, R.; Wachtel, M.S.; Ginsberg, M.D. Induction of reproducible brain infarction by photochemically initiated thrombosis. Ann. Neurol. 1985, 17, 497–504. [Google Scholar] [CrossRef] [PubMed]
  27. Labat-gest, V.; Tomasi, S. Photothrombotic ischemia: A minimally invasive and reproducible photochemical cortical lesion model for mouse stroke studies. J. Vis. Exp. 2013, 76, e50370. [Google Scholar] [CrossRef] [Green Version]
  28. Sun, Y.Y.; Kuo, Y.M.; Chen, H.R.; Short-Miller, J.C.; Smucker, M.R.; Kuan, C.Y. A murine photothrombotic stroke model with an increased fibrin content and improved responses to tPA-lytic treatment. Blood Adv. 2020, 4, 1222–1231. [Google Scholar] [CrossRef]
  29. Sigler, A.; Goroshkov, A.; Murphy, T.H. Hardware and methodology for targeting single brain arterioles for photothrombotic stroke on an upright microscope. J. Neurosci. Methods 2008, 170, 35–44. [Google Scholar] [CrossRef]
  30. Alaverdashvili, M.; Paterson, P.G.; Bradley, M.P. Laser system refinements to reduce variability in infarct size in the rat photothrombotic stroke model. J. Neurosci. Methods 2015, 247, 58–66. [Google Scholar] [CrossRef] [Green Version]
  31. Tuor, U.I.; Deng, Q.; Rushforth, D.; Foniok, T.; Qiao, M. Model of minor stroke with mild peri-infarct ischemic injury. J. Neurosci. Methods 2016, 268, 56–65. [Google Scholar] [CrossRef] [Green Version]
  32. Oh, S.S.; Park, H.J.; Min, H.S.; Kim, S.D.; Bae, S.K.; Kim, J.S.; Ryu, R.-H.; Kim, J.C.; Kim, S.H.; Lee, S.; et al. Investigation of a photothrombosis inducing system for an observation of transient variations in an in vivo rat brain. Curr. Opt. Photonics 2018, 2, 499–507. [Google Scholar]
  33. Kim, Y.; Lee, Y.B.; Bae, S.K.; Oh, S.S.; Choi, J. Development of a photochemical thrombosis investigation system to obtain a rabbit ischemic stroke model. Sci. Rep. 2021, 11, 5787. [Google Scholar] [CrossRef]
  34. Liu, Q.; Li, Y.; Lu, H.; Tong, S. Real-time high resolution laser speckle imaging of cerebral vascular changes in a rodent photothrombosis model. Biomed. Opt. Express 2014, 5, 1483–1493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Zhang, S.; Murphy, T.H. Imaging the impact of cortical microcirculation on synaptic structure and sensory-evoked hemodynamic responses in vivo. PLoS Biol. 2007, 5, e119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Lu, H.; Li, Y.; Yuan, L.; Li, H.; Lu, X.; Tong, S. Induction and imaging of photothrombotic stroke in conscious and freely moving rats. J. Biomed. Opt. 2014, 19, 096013. [Google Scholar] [CrossRef] [PubMed]
  37. Xia, T.; Li, N.; Fang, X. Single-molecule fluorescence imaging in living cells. Ann. Rev. Phys. Chem. 2013, 64, 459–480. [Google Scholar] [CrossRef]
  38. Biswas, J.; Liu, Y.; Singer, R.H.; Wu, B. Fluorescence imaging methods to investigate translation in single cells. Cold Spring Harb. Perspect. Biol. 2019, 11, a032722. [Google Scholar] [CrossRef]
  39. Choi, J.; Sung, J.H.; Shuler, M.L.; Kim, D. Investigation of portable in situ fluorescence optical detection for microfluidic 3D cell culture assays. Opt. Lett. 2016, 35, 1374–1376. [Google Scholar] [CrossRef]
  40. Morris, S.J. Real-time multi-wavelength fluorescence imaging of living cells. Biotechnoques 1990, 8, 296–308. [Google Scholar]
  41. Mitsunaga, M.; Kosaka, N.; Choyke, P.L.; Young, M.R.; Dextras, C.R.; Saud, S.M.; Colburn, N.H.; Sakabe, M.; Nagano, T.; Asanuma, D.; et al. Fluorescence endoscopic detection of murine colitis-associated colon cancer by topically applied enzymatically rapid-activatable probe. Gut 2013, 62, 1179–1186. [Google Scholar] [CrossRef]
  42. Nudo, R.J. Functional and structural plasticity in mortor cortex: Implications for stroke recovery. Phys. Med. Rehabil. Clin. N. Am. 2003, 14, S57–S76. [Google Scholar] [CrossRef]
  43. Murphy, T.H.; Corbett, D. Plasticity during stroke recovery: From synapse to behavior. Nat. Rev. Neurosci. 2009, 10, 861–872. [Google Scholar] [CrossRef]
  44. Okabe, N.; Narita, K.; Miyamoto, O. Axonal remodeling in the corticospinal tract after stroke: How does rehabilitative training modulate it? Neural Regen. Res. 2017, 12, 185–192. [Google Scholar] [CrossRef] [PubMed]
  45. Carmichael, S.T.; Kathirvelu, B.; Schweppe, C.A.; Nie, E.H. Molecular, cellular and functional events in axonal sprouting after stroke. Exp. Neurol. 2017, 287, 384–394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Brown, C.E.; Boyd, J.D.; Murphy, T.H. Longitudinal in vivo imaging reveals balanced and branch-specific remodeling of mature cortical pyramidal dendritic arbors after stroke. J. Cereb. Blood Flow Metabol. 2010, 30, 783–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Nie, S.; Chiu, D.T.; Zare, R.N. Probing individual molecules with confocal fluorescence microscopy. Science 1994, 266, 1018–1021. [Google Scholar] [CrossRef]
  48. Nwaneshiudu, A.; Kuschal, C.; Sakamoto, F.H.; Anderson, R.R.; Schwarzenberger, K.; Young, R.C. Introduction to confocal microscopy. J. Investig. Dermatol. 2012, 132, 1–5. [Google Scholar] [CrossRef] [Green Version]
  49. Helmchen, F.; Denk, W. Deep tissue two-photon microscopy. Nat. Methods 2005, 2, 932–940. [Google Scholar] [CrossRef]
  50. Zong, W.; Wu, R.; Li, M.; Hu, Y.; Li, Y.; Li, J.; Rong, H.; Wu, H.; Xu, Y.; Lu, Y.; et al. Fast high-resolution miniature two-photon microscopy for brain imaging in freely behaving mice. Nat. Methods 2017, 14, 713–719. [Google Scholar] [CrossRef]
  51. Nishimura, N.; Schaffer, C.B.; Friedman, B.; Lyden, P.D.; Kleinfeld, D. Penetrating arterioles are a bottleneck in the perfusion of neocortex. Proc. Natl. Acad. Sci. USA 2007, 104, 365–370. [Google Scholar] [CrossRef] [Green Version]
  52. Watts, L.T.; Zheng, W.; Carling, R.J.; Frohlich, V.C.; Lechleiter, J.D. Rose Bengal photothrombosis by confocal optical imaging in vivo: A model of single vessel stroke. J. Vis. Exp. 2015, 100, e52794. [Google Scholar]
  53. Schrandt, C.J.; Kazmi, S.S.; Jonse, T.A.; Dunn, A.K. Chronic monitoring of vascular progression after ischemic stroke using multiexposure speckle imaging and two-photon fluorescence microscopy. J. Cereb. Blood Flow Metabol. 2015, 35, 933–942. [Google Scholar] [CrossRef]
  54. Aswendt, M.; Adamczak, J.; Tennstaedt, A. A review of novel optical imaging strategies of the stroke pathology and stem cell therapy in stroke. Front. Cell. Neurosci. 2014, 8, 226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Hong, G.; Diao, S.; Chang, J.; Antaris, A.L.; Chen, C.; Zhang, B.; Zhao, S.; Atochin, D.N.; Huang, P.L.; Andreasson, K.I.; et al. Through-skull fluorescence imaging of the brain in a new near-infrared window. Nat. Photon. 2014, 8, 723–730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Zhang, Y.; Fan, S.; Yao, Y.; Ding, J.; Wang, Y.; Zhao, Z.; Liao, L.; Li, P.; Zang, F.; Teng, G.-J. In vivo near-infrared imaging of fibrin deposition in thromboembolic stroke in mice. PLoS ONE 2012, 7, e30262. [Google Scholar] [CrossRef] [PubMed]
  57. Bunce, S.C.; Izzetoglu, M.; Izzetoglu, K.; Onaral, B.; Pourrezaei, K. Functional near-infrared spectroscopy. IEEE Eng. Med. Biol. Mag. 2006, 25, 54–62. [Google Scholar] [CrossRef]
  58. Scholkmann, F.; Kleiser, S.; Metz, A.J.; Zimmermann, R.; Pavia, J.M.; Wolf, U.; Wolf, M. A review on continuous wave functional near-infrared spectroscopy and imaging instrumentation and methodology. Neuroimage 2014, 85, 6–27. [Google Scholar] [CrossRef]
  59. Ferrari, M.; Quaresima, V. A brief review on the history of human functional near-infrared spectroscopy (fNIRS) development and fields of application. Neuroimage 2012, 63, 921–935. [Google Scholar] [CrossRef]
  60. Nguyen, H.D.; Hong, K.S.; Shin, Y.I. Bundled-optode method in functional near-infrared spectroscopy. PLoS ONE 2016, 11, e0165146. [Google Scholar] [CrossRef]
  61. Shang, Y.; Chen, L.; Toborek, M.; Yu, G. Diffuse optical monitoring of repeated cerebral ischemia in mice. Opt. Express 2011, 19, 20301–20315. [Google Scholar] [CrossRef] [Green Version]
  62. Han, C.-H.; Song, H.; Kang, Y.-G.; Kim, B.-M.; Im, C.-H. Hemodynamic responses in rat brain during transcranial direct current stimulation: A functional near-infrared spectroscopy study. Biomed. Opt. Express 2014, 5, 1812–1821. [Google Scholar] [CrossRef] [Green Version]
  63. Lin, Z.-J.; Ren, M.; Li, L.; Liu, Y.; Su, J.; Yang, S.-H.; Liu, H. Interleaved imaging of cerebral hemodynamics and blood flow index to monitor ischemic stroke and treatment in rat by volumetric diffuse optical tomography. Neuroimage 2014, 85, 566–582. [Google Scholar] [CrossRef] [Green Version]
  64. Aries, M.J.H.; Coumou, A.D.; Elting, J.W.J.; van der Harst, J.J.; Kremer, B.P.H.; Vroomen, P.C.A.J. Near infrared spectroscopy for the detection of desaturations in vulnerable ischemic brain tissue: A pilot study at the stroke unit bedside. Stroke 2012, 43, 1134–1136. [Google Scholar] [CrossRef] [PubMed]
  65. Giacalone, G.; Zanoletti, M.; Re, R.; Germinario, B.; Contini, D.; Spinelli, L.; Torricelli, A.; Roveri, L. Time-domain near-infrared spectroscopy in acute ischemic stroke patients. Neurophotonics 2019, 6, 015003. [Google Scholar] [CrossRef] [PubMed]
  66. Saitou, H.; Yanagi, H.; Hara, S.; Tsuchiya, S.; Tomura, S. Cerebral blood volume and oxygenation among poststroke hemiplegic patients: Effects of 13 rehabilitation tasks measured by near-infrared spectroscopy. Arch. Phys. Med. Rehabil. 2000, 81, 1348–1356. [Google Scholar] [CrossRef] [PubMed]
  67. Kato, H.; Izumiyama, M.; Koizumi, H.; Takahashi, A.; Itoyama, Y. Near-infrared spectroscopic topography as a tool to monitor motor reorganization after hemiparetic stroke: A comparison with functional MRI. Stroke 2002, 33, 2032–2036. [Google Scholar] [CrossRef] [PubMed]
  68. Lin, P.Y.; Lin, S.I.; Penney, T.; Chen, J.J. Applications of near infrared spectroscopy and imaging for motor rehabilitation in stroke patients. J. Med. Biol. Eng. 2009, 29, 210–221. [Google Scholar]
  69. Miyai, I.; Suzuki, M.; Hatakenaka, M.; Kubota, K. Effect of body weight support on cortical activation during gait in patients with stroke. Exp. Brain Res. 2006, 169, 85–91. [Google Scholar] [CrossRef]
  70. Miyai, I.; Yagura, H.; Oda, I.; Konishi, I.; Eda, H.; Suzuki, T.; Kubota, K. Premotor cortex is involved in restoration of gait in stroke. Ann. Neurol. 2002, 52, 188–194. [Google Scholar] [CrossRef]
  71. Lin, P.-Y.; Chen, J.-J.J.; Lin, S.-I. The cortical control of cycling exercise in stroke patients: An fNIRS study. Hum. Brain Map. 2013, 34, 2381–2390. [Google Scholar] [CrossRef]
  72. Dutta, A.; Jacob, A.; Chowdhury, S.R.; Das, A.; Nitsche, M.A. EEG-NIRS based assessment of neurovascular coupling during anodal transcranial direct current stimulation—A stroke case series. J. Med. Syst. 2015, 39, 36. [Google Scholar] [CrossRef]
  73. Guhathakurta, D.; Dutta, A. Computational pipeline for NIRS-EEG joint imaging of tDCS-evoked cerebral responses—An application in ischemic stroke. Front. Neurosci. 2016, 10, 261. [Google Scholar] [CrossRef] [Green Version]
  74. Litscher, G.; Zhang, X.; Sheng, Z.; Jing, X.-H.; Wang, L. Multimodal laser stimulation and traditional needle acupuncture in post-stroke patients—A pilot cross-over study with results from near infrared spectroscopy. Medicines 2019, 6, 115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Kim, A.L. Portable traumatic brain injury detection with near-infrared technology: Infrascanner model 2000. Mil. Med. 2015, 180, 597–598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Peters, J.; Van Wageningen, B.; Hoogerwerf, N.; Tan, E. Near-infrared spectroscopy: A promising prehospital tool for management of traumatic brain injury. Prehosp. Disaster Med. 2017, 32, 414–418. [Google Scholar] [CrossRef] [PubMed]
  77. Schober, P.; Bossers, S.M.; Schwarte, L.A. Intracranial hematoma detection by near infrared spectroscopy in a helicopter emergency medical service: Practical experience. BioMed Res. Int. 2017, 2017, 1846830. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Fujimoto, J.G.; Brezinski, M.E.; Tearney, G.J.; Boppart, S.A.; Bouma, B.; Hee, M.R.; Southern, J.F.; Swanson, E.A. Optical biopsy and imaging using optical coherence tomography. Nat. Med. 1995, 1, 970–972. [Google Scholar] [CrossRef]
  79. Fercher, A.F. Optical coherence tomography. J. Biomed. Opt. 1996, 1, 157–174. [Google Scholar] [CrossRef]
  80. Fercher, A.F.; Drexler, W.; Hitzenberger, C.K.; Lasser, T. Optical coherence tomography-principles and applications. Rep. Prog. Phys. 2003, 66, 239. [Google Scholar] [CrossRef]
  81. Wojtkowski, M.; Leitgeb, R.; Kowalczyk, A.; Bajraszewski, T.; Fercher, A.F. In vivo human retinal imaging by Fourier domain optical coherence tomography. J. Biomed. Opt. 2002, 7, 457–463. [Google Scholar] [CrossRef]
  82. Graham, K.L.; McCowan, C.I.; Caruso, K.; Billson, F.M.; Whittaker, C.J.; White, A. Optical coherence tomography of the retina, nerve fiber layer, and optic nerve head in dogs with glaucoma. Vet. Ophthalmol. 2020, 23, 97–112. [Google Scholar] [CrossRef]
  83. Rajabi-Estarabadi, A.; Bittar, J.M.; Zheng, C.; Nascimento, V.; Camacho, I.; Feun, L.G.; Nasiriavanaki, M.; Kunz, M.; Nouri, K. Optical coherence tomography imaging of melanoma skin cancer. Lasers Med. Sci. 2019, 34, 411–420. [Google Scholar] [CrossRef]
  84. Rangaraju, L.P.; Kunapuli, G.; Every, D.; Ayala, O.D.; Ganapathy, P.; Mahadevan-Jansen, A. Classification of burn injury using Raman spectroscopy and optical coherence tomography: An ex-vivo study on porcine skin. Burns 2019, 45, 659–670. [Google Scholar] [CrossRef] [PubMed]
  85. Wartak, A.; Beer, F.; Desissaire, S.; Baumann, B.; Pircher, M.; Hitzenberger, C.K. Investigating spontaneous retinal venous pulsation using Doppler optical coherence tomography. Sci. Rep. 2019, 9, 4237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Israelsen, N.M.; Peterson, C.R.; Barh, A.; Jain, D.; Jensen, M.; Hannesschläger, G.; Tidemand-Lichtenberg, P.; Pedersen, C.; Podoleanu, A.; Bang, O. Real-time high-resolution mid-infrared optical coherence tomography. Light Sci. Appl. 2019, 8, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Choi, W.J.; Li, Y.; Wang, R.K. Monitoring acute stroke progression: Multi-parametric OCT imaging of cortical perfusion, flow, and tissue scattering in a mouse model of permanent focal ischemia. IEEE Trans. Med. Imaging 2019, 38, 1427–1437. [Google Scholar] [CrossRef]
  88. Wei, X.; Hormel, T.T.; Pi, S.; Guo, Y.; Jian, Y.; Jia, Y. High dynamic range optical coherence tomography angiography (HDR-OCTA). Biomed. Opt. Express 2019, 10, 3560–3571. [Google Scholar] [CrossRef]
  89. Yu, L.; Nguyen, E.; Liu, G.; Choi, B.; Chen, Z. Spectral Doppler optical coherence tomography imaging of localized ischemic stroke in a mouse model. J. Biomed. Opt. 2010, 15, 066006. [Google Scholar] [CrossRef] [Green Version]
  90. Osiac, E.; Bălşeanu, T.A.; Mogoantă, L.; Gheonea, D.I.; Pirici, I.; Iancău, M.; Mitran, S.I.; Albu, C.V.; Cătălin, B.; Sfredel, V. Optical coherence tomography investigation of ischemic stroke inside a rodent model. Rom. J. Morphol. Embryol. 2014, 55, 767–772. [Google Scholar]
  91. Streba, C.T.; Georgescu, S.L.; Jigau, M.; Dinu, M.A.; Balseanu, T.-A.; Dinescu, V.C.; Catalin, B.; Osiac, E. Artificial neural network as an analyze tool for optical coherence tomography images of experimental stroke models-a pilot study. Phys. AUC 2015, 25, 1–8. [Google Scholar]
  92. Srinivasan, V.J.; Mandeville, E.T.; Can, A.; Blasi, F.; Climov, M.; Daneshmand, A.; Lee, J.H.; Yu, E.; Radhakrishnan, H.; Lo, E.H.; et al. Multiparametric, longitudinal optical coherence tomography imaging reveals acute injury and chronic recovery in experimental ischemic stroke. PLoS ONE 2013, 8, e71478. [Google Scholar] [CrossRef]
  93. Chen, S.; Liu, Q.; Shu, X.; Soetikno, B.; Tong, S.; Zhang, H.F. Imaging hemodynamic response after ischemic stroke in mouse cortex using visible-light optical coherence tomography. Biomed. Opt. Express 2016, 7, 3377–3389. [Google Scholar] [CrossRef] [Green Version]
  94. Liu, Q.; Chen, S.; Soetikno, B.; Liu, W.; Tong, S.; Zhang, H.F. Monitoring acute stroke in mouse model using laser speckle imaging-guided visible-light optical coherence tomography. IEEE Trans. Biomed. Eng. 2017, 65, 2136–2142. [Google Scholar] [CrossRef] [PubMed]
  95. Xu, M.; Wang, L.V. Photoacoustic imaging in biomedicine. Rev. Sci. Instrum. 2006, 77, 041101. [Google Scholar] [CrossRef] [Green Version]
  96. Beard, P. Biomedical photoacoustic imaging. Interface Focus 2011, 1, 602–631. [Google Scholar] [CrossRef] [PubMed]
  97. Li, L.; Lin, L.; Wang, L.V. Multiscale photoacoustic tomography. Opt. Photonics News 2018, 29, 32–39. [Google Scholar] [CrossRef]
  98. Schwarz, M.; Omar, M.; Buehler, A.; Aguirre, J.; Ntziachristos, V. Implications of ultrasound frequency in optoacoustic mesoscopy of the skin. IEEE. Trans. Med. Imaging 2015, 34, 672–677. [Google Scholar] [CrossRef]
  99. Fehm, T.F.; Deán-Ben, X.L.; Ford, S.J.; Razansky, D. In vivo whole-body optoacoustic scanner with real-time volumetric imaging capacity. Optica 2016, 3, 1153–1159. [Google Scholar] [CrossRef]
  100. Li, L.; Zhu, L.; Ma, C.; Lin, L.; Yao, J.; Wang, L.; Maslov, K.; Zhang, R.; Chen, W.; Shi, J.; et al. Single-impulse panoramic photoacoustic computed tomography of small-animal whole-body dynamics at high spatiotemporal resolution. Nat. Biomed. Eng. 2017, 1, 0071. [Google Scholar] [CrossRef]
  101. Hartman, R.K.; Hallam, K.A.; Donnelly, E.M.; Emelianov, S.Y. Photoacoustic imaging of gold nanorods in the brain delivered via microbubble-assisted focused ultrasound: A tool for in vivo molecular neuroimaging. Laser Phys. Lett. 2019, 16, 025603. [Google Scholar] [CrossRef]
  102. Kitai, T.; Torii, M.; Sugie, T.; Kanao, S.; Mikami, Y.; Shiina, T.; Toi, M. Photoacoustic mammography: Initial clinical results. Breast Cancer 2014, 21, 146–153. [Google Scholar] [CrossRef] [Green Version]
  103. Lin, L.; Hu, P.; Shi, J.; Appleton, C.M.; Maslov, K.; Li, L.; Zhang, R.; Wang, L.V. Single-breath-hold photoacoustic computed tomography of the breast. Nat. Commun. 2018, 9, 2352. [Google Scholar] [CrossRef]
  104. Laufer, J.; Zhang, E.; Raivich, G.; Beard, P. Three-dimensional noninvasive imaging of the vasculature in the mouse brain using a high resolution photoacoustic scanner. Appl. Opt. 2009, 48, D299–D306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Lin, L.; Yao, J.; Zhang, R.; Chen, C.-C.; Huang, C.-H.; Li, Y.; Wang, L.; Chapman, W.; Zou, J.; Wang, L.V. High-speed photoacoustic microscopy of mouse cortical microhemodynamics. J. Biophotonics 2017, 10, 792–798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Kneipp, M.; Turner, J.; Hambauer, S.; Krieg, S.M.; Lehmberg, J.; Lindauer, U.; Razansky, D. Functional real-time optoacoustic imaging of middle cerebral artery occlusion in mice. PLoS ONE 2014, 9, e96118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Deng, Z.; Wang, Z.; Yang, X.; Luo, Q.; Gong, H. In vivo imaging of hemodynamics and oxygen metabolism in acute focal cerebral ischemic rats with laser speckle imaging and functional photoacoustic microscopy. J. Biomed. Opt. 2012, 17, 081415. [Google Scholar] [CrossRef]
  108. Wu, D.; Yang, J.; Zhang, G.; Jiang, H. Noninvasive in vivo monitoring of collagenase induced intracerebral hemorrhage by photoacoustic tomography. Biomed. Opt. Express 2017, 8, 2276–2286. [Google Scholar] [CrossRef] [Green Version]
  109. Ni, R.; Vaas, M.; Ren, W.; Klohs, J. Non-invasive detection of matrix-metalloproteinase activity in a mouse model of cerebral ischemia using multispectraloptoacoustic tomography. Proc. SPIE 2018, 10494, 104940O. [Google Scholar]
  110. Medrado, A.P.; Soares, A.P.; Santos, E.T.; Reis, S.R.A.; Andrade, Z.A. Influence of laser photobiomodulation upon connective tissue remodeling during wound healing. J. Photochem. Photobiol. B 2008, 92, 144–152. [Google Scholar] [CrossRef]
  111. Kuffler, D.P. Photobiomodulation in promoting wound healing: A review. Future Med. 2016, 11, 107–122. [Google Scholar] [CrossRef]
  112. Corazza, A.V.; Jorge, J.; Kurachi, C.; Bagnato, V.S. Photobiomodulation on the angiogenesis of skin wounds in rats using different light sources. Photomed. Laser Surg. 2007, 25, 102–106. [Google Scholar] [CrossRef]
  113. de Freitas, L.F.; Hamblin, M.R. Proposed mechanisms of photobiomodulation or low-level light therapy. IEEE J. Sel. Top. Quantum Electron. 2016, 22, 348–364. [Google Scholar] [CrossRef] [Green Version]
  114. Desmet, K.D.; Paz, D.A.; Corry, J.J.; Eells, J.T.; Wong-Riley, M.T.T.; Henry, M.M.; Buchmann, E.V.; Connelly, M.P.; Dovi, J.V.; Liang, H.L.; et al. Clinical and experimental applications of NIR-LED photobiomodulation. Photomed. Laser Surg. 2006, 24, 121–128. [Google Scholar] [CrossRef] [PubMed]
  115. Cassano, P.; Petrie, S.R.; Hamblin, M.R.; Henderson, T.A.; Iosifescu, D.V. Review of transcranial photobiomodulation for major depressive disorder: Targeting brain metabolism, inflammation, oxidative stress, and neurogenesis. Neurophotonics 2016, 3, 031404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Cassano, P.; Petrie, S.R.; Mischoulon, D.; Cusin, C.; Katnani, H.; Yeung, A.; Taboada, L.D.; Archibald, A.; Bui, E.; Baer, L.; et al. Transcranial photobiomodulation for the treatment of major depressive disorder. The ELATED-2 pilot trial. Photomed. Laser Surg. 2018, 36, 634–646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Hamilton, C.; Hamilton, D.; Nicklason, F.; El Massri, N.; Mitrofanis, J. Exploring the use of transcranial photobiomodulation in Parkinson’s disease patients. Neural Regen. Res 2018, 13, 1738. [Google Scholar]
  118. Hamblin, M.R. Shining light on the head: Photobiomodulation for brain disorders. BBA Clin. 2016, 6, 113–124. [Google Scholar] [CrossRef] [Green Version]
  119. DeTaboada, L.; Ilic, S.; Leichliter-Martha, S.; Oron, U.; Oron, A.; Streeter, J. Transcranial application of low-energy laser irradiation improves neurological deficits in rats following acute stroke. Laser Surg. Med. 2006, 38, 70–73. [Google Scholar] [CrossRef]
  120. Oron, A.; Oron, U.; Chen, J.; Eilam, A.; Zhang, C.; Sadeh, M.; Lampl, Y.; Streeter, J.; DeTaboada, L.; Chopp, M. Low-level laser therapy applied transcranially to rats after induction of stroke significantly reduces long-term neurological deficits. Stroke 2006, 37, 2620–2624. [Google Scholar] [CrossRef] [Green Version]
  121. Lapchak, P.A.; Salgado, K.F.; Chao, C.H.; Zivin, J.A. Transcranial near-infrared light therapy improves motor function following embolic strokes in rabbits: An extended therapeutic window study using continuous and pulse frequency delivery modes. Neuroscience 2007, 148, 907–914. [Google Scholar] [CrossRef]
  122. Lee, H.I.; Lee, S.-W.; Kim, N.G.; Park, K.-J.; Choi, B.T.; Shin, Y.-I.; Shin, H.K. Low-level light emitting diode (LED) therapy suppresses inflammasome-mediated brain damage in experimental ischemic stroke. J. Biophotonics 2017, 11, 1502–1513. [Google Scholar] [CrossRef]
  123. Lee, H.I.; Lee, S.W.; Kim, N.G.; Park, K.-J.; Choi, B.T.; Shin, Y.-I.; Shin, H.K. Low-level light emitting diode therapy promotes long-term functional recovery after experimental stroke in mice. J. Biophotonics 2017, 10, 1761–1771. [Google Scholar] [CrossRef]
  124. Argibay, B.; Compos, F.; Perez-Mato, M.; Vieites-Prado, A.; Correa-Paz, C.; López-Arias, E.; Da Silva-Candal, A.; Moreno, V.; Montero, C.; Sobrino, T.; et al. Light-Emitting Diode photobiomodulation after cerebral ischemia. Front. Neurol. 2019, 10, 911. [Google Scholar] [CrossRef] [PubMed]
  125. Lampl, Y.; Zivin, J.A.; Fisher, M.; Lew, R.; Welin, L.; Dahlof, B.; Borenstein, P.; Andersson, B.; Perez, J.; Caparo, C.; et al. Infrared laser therapy for ischemic stroke: A new treatment strategy: Results of the NeuroThera Effectiveness and Safety Trial-1 (NEST-1). Stroke 2007, 38, 1843–1849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Lapchak, P. Taking a light approach to treating acute ischemic stroke patients: Transcranial near-infrared laser therapy translational science. Ann. Med. 2010, 42, 576–586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Huisa, B.N.; Stemer, A.B.; Walker, M.G.; Rapp, K.; Meyer, B.C.; Zivin, J.A. Transcranial laser therapy for acute ischemic stroke: A pooled analysis of NEST-1 and NEST-2. Int. J. Stroke 2013, 8, 315–320. [Google Scholar] [CrossRef] [Green Version]
  128. Casalechi, H.L.; Dumont, A.J.L.; Ferreira, L.A.B.; de Paiva, P.R.V.; Machado, C.D.S.M.; de Carvalho, P.D.T.C.; Oliveira, C.S.; Leal, E.C.P., Jr. Acute effects of photobiomodulation therapy and magnetic field on functional mobility in stroke survivors: A randomized, sham-controlled, triple-blind, crossover, clinical trial. Lasers Med. Sci. 2019, 35, 1253–1262. [Google Scholar] [CrossRef] [PubMed]
  129. Nagel, G.; Szellas, T.; Huhn, W.; Kateriya, S.; Adeishvili, N.; Berthold, P.; Ollig, D.; Hegemann, P.; Bamberg, E. Channelrhodopsin-2, a directly light-gated cation-selective membrane channel. Proc. Natl. Acad. Sci. USA 2003, 100, 13940–13945. [Google Scholar] [CrossRef] [Green Version]
  130. Kim, C.K.; Adhikari, A.; Deisseroth, K. Integration of optogenetics with complementary methodologies in systems neuroscience. Nat. Rev. Neurosci. 2017, 18, 222–235. [Google Scholar] [CrossRef]
  131. Song, C.; Knöpfel, T. Optogenetics enlightens neuroscience drug discovery. Nat. Rev. Drug Discov. 2016, 15, 97–109. [Google Scholar] [CrossRef]
  132. Cheng, M.Y.; Wang, E.H.; Woodson, W.J.; Wang, S.; Sun, G.; Lee, A.G.; Arac, A.; Fenno, L.E.; Deisseroth, K.; Steinberg, G.K. Optogenetic neuronal stimulation promotes functional recovery after stroke. Proc. Natl. Acad. Sci. USA 2014, 111, 12913–12918. [Google Scholar] [CrossRef] [Green Version]
  133. Song, M.; Yu, S.P.; Mohamad, O.; Cao, W.; Wei, Z.Z.; Gu, X.; Jiang, M.Q.; Wei, L. Optogenetic stimulation of glutamatergic neuronal activity in the striatum enhances neurogenesis in the subventricular zone of normal and stroke mice. Neurobiol. Dis. 2017, 98, 9–24. [Google Scholar] [CrossRef] [Green Version]
  134. Shah, A.M.; Ishizaka, S.; Cheng, M.Y.; Wang, E.H.; Bautista, A.R.; Levy, S.; Smerin, D.; Sung, G.; Steinberg, G.K. Optogenetic neuronal stimulation of the lateral cerebellar nucleus promotes persistent functional recovery after stroke. Sci. Rep. 2017, 7, 46612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Pendharkar, A.V.; Smerin, D.; Gonzalez, L.; Wang, E.H.; Levy, S.; Wang, S.; Ishizaka, S.; Ito, M.; Uchino, H.; Chiang, T.; et al. Optogenetic stimulation reduces neuronal nitric oxide synthase expression after stroke. Transl. Stroke Res. 2021, 12, 347–356. [Google Scholar] [CrossRef] [PubMed]
  136. Conti, E.; Mascaro, A.L.A.; Scaglione, A.; de Vito, G.; Calugi, F.; Pasquini, M.; Pizzorusso, T.; Micera, S.; Pavone, F.S. Restoration of motor-evoked cortical activity is a distinguishing feature of the most effective rehabilitation therapy after stroke. bioRxiv 2021. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) Generation of photothrombotic stroke on a single brain arteriole using a photothrombosis inducing module, which was employed using a conventional upright microscope. Images acquired while blocking a single arteriole through photochemical thrombosis and light irradiation for 6, 20, and 120 s are depicted on the right. Reprinted with permission from Ref. [29]. Copyright 2008 Elsevier B.V. (b) A fiber-optic light irradiation probe used to induce photochemical thrombosis in a minor stroke, and magnetic resonance T2- and diffusion-weighted images of a brain lesion confirmed that the investigated probe can induce minor stroke. Reprinted with permission from Ref. [31]. Copyright 2016 Elsevier B.V.
Figure 1. (a) Generation of photothrombotic stroke on a single brain arteriole using a photothrombosis inducing module, which was employed using a conventional upright microscope. Images acquired while blocking a single arteriole through photochemical thrombosis and light irradiation for 6, 20, and 120 s are depicted on the right. Reprinted with permission from Ref. [29]. Copyright 2008 Elsevier B.V. (b) A fiber-optic light irradiation probe used to induce photochemical thrombosis in a minor stroke, and magnetic resonance T2- and diffusion-weighted images of a brain lesion confirmed that the investigated probe can induce minor stroke. Reprinted with permission from Ref. [31]. Copyright 2016 Elsevier B.V.
Applsci 12 01891 g001
Figure 2. Two-photon high-resolution fluorescence images used to observe variations in cerebral blood micro-vessels and dendrites before and after 30 min of stroke induction in a large area (>0.40 mm2). They indicate that dendrites do not recover from stroke and its consequent brain damages in a large area. Reprinted with permission from Ref. [35]. Copyright 2007 Public Library of Science.
Figure 2. Two-photon high-resolution fluorescence images used to observe variations in cerebral blood micro-vessels and dendrites before and after 30 min of stroke induction in a large area (>0.40 mm2). They indicate that dendrites do not recover from stroke and its consequent brain damages in a large area. Reprinted with permission from Ref. [35]. Copyright 2007 Public Library of Science.
Applsci 12 01891 g002
Figure 4. Schematic of an optical coherent tomographic imaging system for in vivo brain angiography and optical coherence tomography angiographic images of acute ischemic mice before inducing stroke and after one week. Reprinted with permission from Ref. [92]. Copyright 2013 Public Library of Science.
Figure 4. Schematic of an optical coherent tomographic imaging system for in vivo brain angiography and optical coherence tomography angiographic images of acute ischemic mice before inducing stroke and after one week. Reprinted with permission from Ref. [92]. Copyright 2013 Public Library of Science.
Applsci 12 01891 g004
Figure 5. Deoxy-hemoglobin distributions, which were estimated using multispectral photoacoustic brain images in mice with ischemic stroke through middle cerebral artery occlusion. The figure below describes ex vivo brain slices for verification of stroke induction and their comparison with photoacoustic images. Reprinted with permission from Ref. [106]. Copyright 2014 Public Library of Science.
Figure 5. Deoxy-hemoglobin distributions, which were estimated using multispectral photoacoustic brain images in mice with ischemic stroke through middle cerebral artery occlusion. The figure below describes ex vivo brain slices for verification of stroke induction and their comparison with photoacoustic images. Reprinted with permission from Ref. [106]. Copyright 2014 Public Library of Science.
Applsci 12 01891 g005
Figure 6. Therapeutic effects of three different conditions (continuous wave (CW) and pulsed light (PULSE 1 and PULSE 2)) of infrared light irradiation treatments six hours after an induction of acute ischemic stroke in rabbits. Pulsed light irradiation treatments in conditions of both PULSE 1 and PULSE 2 have shown statistically significant therapeutic effects, while treatments in conditions of CW did not. Reprinted with permission from Ref. [121]. Copyright 2007 Elsevier B.V.
Figure 6. Therapeutic effects of three different conditions (continuous wave (CW) and pulsed light (PULSE 1 and PULSE 2)) of infrared light irradiation treatments six hours after an induction of acute ischemic stroke in rabbits. Pulsed light irradiation treatments in conditions of both PULSE 1 and PULSE 2 have shown statistically significant therapeutic effects, while treatments in conditions of CW did not. Reprinted with permission from Ref. [121]. Copyright 2007 Elsevier B.V.
Applsci 12 01891 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Oh, S.S.; Kim, Y.; Lee, Y.B.; Bae, S.K.; Kim, J.S.; An, S.-h.; Choi, J.-r. Optical Modalities for Research, Diagnosis, and Treatment of Stroke and the Consequent Brain Injuries. Appl. Sci. 2022, 12, 1891. https://doi.org/10.3390/app12041891

AMA Style

Oh SS, Kim Y, Lee YB, Bae SK, Kim JS, An S-h, Choi J-r. Optical Modalities for Research, Diagnosis, and Treatment of Stroke and the Consequent Brain Injuries. Applied Sciences. 2022; 12(4):1891. https://doi.org/10.3390/app12041891

Chicago/Turabian Style

Oh, Sung Suk, Yoonhee Kim, Yoon Bum Lee, Seung Kuk Bae, Jun Sik Kim, Sang-hyun An, and Jong-ryul Choi. 2022. "Optical Modalities for Research, Diagnosis, and Treatment of Stroke and the Consequent Brain Injuries" Applied Sciences 12, no. 4: 1891. https://doi.org/10.3390/app12041891

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

Article Metrics

Back to TopTop