Next Article in Journal
Triple-Attention Mixed-Link Network for Single-Image Super-Resolution
Next Article in Special Issue
Ultrahigh Resolution Polarization Sensitive Optical Coherence Tomography of the Human Cornea with Conical Scanning Pattern and Variable Dispersion Compensation
Previous Article in Journal
Mechanisms of Control Authority by Nanosecond Pulsed Dielectric Barrier Discharge Actuators on Flow Separation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of OCT in the Gastrointestinal Tract

by
Nicholas S. Samel
1 and
Hiroshi Mashimo
2,*
1
Dartmouth College, Hanover, NH 02090, USA
2
VA Boston Healthcare System, Harvard Medical School, West Roxbury, MA 02132, USA
*
Author to whom correspondence should be addressed.
Appl. Sci. 2019, 9(15), 2991; https://doi.org/10.3390/app9152991
Submission received: 12 July 2019 / Revised: 21 July 2019 / Accepted: 22 July 2019 / Published: 25 July 2019
(This article belongs to the Special Issue Optical Coherence Tomography and Its Applications II)

Abstract

:
Optical coherence tomography (OCT) is uniquely poised for advanced imaging in the gastrointestinal (GI) tract as it allows real-time, subsurface and wide-field evaluation at near-microscopic resolution, which may improve the current limitations or even obviate the need of superficial random biopsies in the surveillance of early neoplasias in the near future. OCT’s greatest impact so far in the GI tract has been in the study of the tubular esophagus owing to its accessibility, less bends and folds and allowance of balloon employment with optimal contact to aid circumferential imaging. Moreover, given the alarming rise in the incidence of Barrett’s esophagus and its progression to adenocarcinoma in the U.S., OCT has helped identify pathological features that may guide future therapy and follow-up strategy. This review will explore the current uses of OCT in the gastrointestinal tract and future directions, particularly with non-endoscopic office-based capsule OCT and the use of artificial intelligence to aid in diagnoses.

1. Introduction

Diagnostic imaging of the gastrointestinal (GI) tract has long relied on white-light endoscopy (WLE) for the detection of mucosal abnormalities such as neoplasias, vascular deformities, and inflammatory conditions. While WLE is a clinical standard for GI imaging and cancer detection, it remains far from ideal. A standard colonoscopy can miss up to 40% of adenomas [1,2] and despite improvements to increase mucosal visualization [3], miss rates are still concerning, especially for flat lesions such as sessile serrated adenomas [4]. While part of the suboptimal detection can be attributed to patients’ inadequate preparatory colonic lavage, WLE’s limited field of view (FOV) and time-consuming mucosal inspection during endoscope withdrawal are additional limiting factors [3]. Since the survival rate from GI cancers can be improved by 5 to 10-fold if diagnosed at an early stage [5], there is an obvious need to improve detection of early GI neoplasias.
Advanced imaging in the GI tract can be broadly divided into surface and subsurface modalities. Surface imaging includes chromoendoscopy [6], virtual chromoendoscopy [7], magnification endoscopy and endocytoscopy [8], and autofluorescence imaging [9]. There have been tremendous improvements in the resolution, field of view (FOV), and methods of surface imaging, such as improving contrast in magnification endoscopy and using selected wavelengths (Narrow Band Imaging, NBI) or digital enhancements such as flexible spectral imaging color enhancement (FICE; Fujinon Inc, Saitama, Japan). However, these modalities are limited to superficial lesions and improved magnification generally comes at the cost of restricted FOV, which makes screening and surveillance for multi-focal lesions more difficult.
On the other hand, subsurface imaging, which includes endoscopic ultrasonography [10], confocal endomicroscopy [11], and OCT, allows for visualization of structures not feasible with most endoscopes. Of these technologies, OCT is a uniquely poised imaging modality, combining nearly microscopic resolution with volumetric and subsurface real-time imaging capabilities [12]. These characteristics position OCT to solve many of the weaknesses of the other GI tract imaging techniques. This review will explore some of the recent advances of OCT in GI applications and describe the future potential for OCT to incorporate purpose-build probe designs and computer-aided diagnosis techniques [12].

2. Evolving OCT Technology

Briefly, the working principle of OCT is analogous to that of an ultrasound, with analysis of reflected light rather than sound waves. An optical ranging technique called low coherence interferometry [13,14] uses a Michelson interferometer to measure the time delay and subsequent signal intensity of light reflected from a tissue sample. Since its first publication in the early 1990s [15], endoscopic OCT techniques [16] and applications have become major areas of research [17,18,19,20].
OCT can be broadly divided into two types: time-domain OCT (TD-OCT) and Fourier-domain OCT (FD-OCT). They differ in that TD-OCT considers each scan individually with a moving mirror (Figure 1a) [21,22,23], while FD-OCT applies a Fourier transform on the detected spectrum (Figure 1b) [24,25,26]. FD-OCT’s introduction in the early 2000s generated a great increase in imaging speed and sensitivity of OCT [23,24,25,26]. FD-OCT can be further divided into subcategories, spectral-domain OCT (SD-OCT) and swept-source OCT (SS-OCT) [21]. SD-OCT uses a spectrometer with a broadband light source and typically operates at shorter wavelengths [27,28]. SS-OCT measures interference spectra using a wavelength-swept light source and a photodetector, operating at longer wavelengths and with a high imaging frequency, sometimes reaching the MHz region [29,30,31]. Due to its improved tissue depth penetration at longer wavelengths, SS-OCT is the technology of choice for endoscopic applications [32,33].
Different imaging probe designs allow different perspectives for acquisition of OCT images. Side-viewing probes are the most common in endoscopic OCT applications due to their large luminal surface area coverage [12]. Different forms of side-viewing probe delivery include the small diameter flexible sheath [16], large diameter inflatable balloon [34] and rigid housing [35]. An introductory video demonstrating imaging in the gastrointestinal tract using a probe-based OCT is available [36]. Small diameter flexible sheath, also called “optical biopsy,” is excellent for obtaining high-speed visualization, but is less resolved than forward viewing probes [16]. The balloon catheter is excellent for tubular structures such as the esophagus, but is not optimized for the bends and folds of the rest of the GI tract. Rigid housing, such as in a capsule, is advantageous for a number of reasons that will be discussed in a subsequent section.
Forward viewing (en face) probes have smaller FOVs but provide a more intuitive viewing scheme akin to magnified endoscopy and endomicroscopy, and are generally integrated with other modes of microscopy and applied to tasks requiring high magnification. Forward-imaging probe designs include microelectromechanical system (MEMS) scanners, piezoelectric transducer (PZT) actuators, and the paired-angle-rotation scanner (PARS) [37,38]. While MEMS scanners have been shown to combat distortion, they often use a nonlinear scan pattern which necessitates oversampling; however, a promising study by Cogliati et al. reported the implementation of pre-shaped open-loop input signals with non-linear parts to achieve live, distortion-free imaging without post-processing with a Gabor-domain optical coherence microscope [39]. Additionally, a MEMS platform has been developed for circumferential scan by Mu et al. using an electrothermal chevron-beam actuator [40]. However, a challenge of MEMS-based scanning endoscopes has been the large probe diameter (4-5.8 mm) [38]. PZT actuators are desirable for their small size and compatibility with an endoscopic accessory channel; however, the length of the tubular PZT actuator is quite long (around 3.5 cm) [37]. Notably, Zhang et al. proposed a reverse-mount PZT design with a shorter rigid length that had comparable performance to forward-mount probes, although performed on colonic samples ex vivo [37]. PARS, which rotates a pair of angle-polished gradient index (GRIN) lenses, is advantageous for its miniaturization and variety of scan modes [41], but retaining concentricity and stable separation of the two GRIN lenses is challenging [38]. Most recently, micromotor catheters have been used to achieve high-speed, volumetric en face ultra-high resolution for the in vivo assessment of BE and dysplasia (Figure 2) [42].
The beam can also be scanned in different patterns such as raster [43], spiral [44] or Lissajou [45]. The use of these scanning patterns depends on the actuators used in the probe design [12].
OCT technology can also be combined with other imaging techniques to give it enhanced molecular sensitivity and imaging depth [38]. However, such endoscopic multi-modal imaging platforms are limited by the small size of the accessory channel of a conventional endoscope [38]. In one paper, such geometric challenges were overcome by putting the imaging and spectroscopy probes together in the same 2mm diameter silica tube [46]. Notwithstanding the challenges of miniaturizing the imaging catheters, OCT has been combined with a diffuse fluorescence spectroscopy probe to obtain simultaneous 1D laser-induced fluorescence (LIF) and 3D OCT spectra in mice. The LIF spectrum allows chemical specificity to the OCT platform [47]. However, multimodal OCT and fluorescence imaging in the GI tract is challenging, as it requires a high speed, high spatial resolution, uniform scanning system with an FDA approved contrast agent [48]. Nevertheless, one study found that layered architecture and microvasculature could be clearly identified using such a multimodal system in the rat rectum [48]. OCT can also be used simultaneously with ultrasound to result in a greater combined image depth [38]. A study in 2010 used a high-frequency ring transducer to focus the ultrasound (US) beam and OCT light beam toward a common imaging spot for simultaneously co-registered US and OCT images to access tissue beyond the traditional 1-2.5 mm OCT imaging depth [38,49]. Additionally, trimodal OCT, US and photoacoustic imaging (PAI) endoscopy have been shown to combine the high resolution subsurface imaging capabilities of OCT, deeper penetration by US and the increased optical absorption contrast of PAI for enhanced visualization of neoplasia [50]. Lastly, polarization sensitive (PS) OCT exploits the altered reflected polarization of light from different materials and tissues to quantitate and provide further contrast [51]. This can be used to detect tumors, such as in the colon, where precancerous lesions have been found to be less depolarizing than benign tissues [51,52]. OCT has been applied to both the upper and lower GI tract either through the accessory channel of an endoscope or independently. While OCT has great potential throughout the entire GI tract, its principal advancements so far have been made in the esophagus.

3. OCT Applications in the GI Tract

3.1. Esophagus

Barrett’s Esophagus (BE) is the main precursor to esophageal adenocarcinoma (EAC). EAC is the eighth most common cancer type in the world, and despite falling incidence rates of squamous cell carcinoma, the incidence of EAC continues to rise globally, especially in western countries (--including the United States) [53,54]. Alarmingly, its incidence increased 6.5-fold from 1975-2009 according to SEER data [55], making the incidence for EAC one of the fastest rising of all cancers in the US, particularly in the white male population [56]. EAC is thought to arise stepwise from specialized intestinal metaplasia (SIM) to low grade dysplasia (LGD), high grade dysplasia (HGD) and then intramucosal carcinoma (IMC) [57,58]. Guidance for surveillance of BE neoplastic progression has been provided by the Seattle Protocol, which advocates for targeted biopsies or resection of suspicious esophageal areas followed by random four-quadrant biopsies at set intervals [59]. This sampling limits itself to less than 5% of the esophageal surface area [12]. Pinch biopsies often do not acquire deeper structures and one study showed they miss the lamina propria more than 60% of the time [60,61], raising concerns regarding under-sampling for the often focal and small areas of dysplasia. Additionally, after ablation, residual BE tissues may become buried under the neosquamous epithelium as subsquamous intestinal metaplasia (SSIM) and elude detection [62].
Given these needs and the relative simplicity of a tubular structure for OCT imaging, OCT in the GI tract has been applied largely to the esophagus. OCT, with its volumetric, subsurface and near-microscopic imaging, is positioned to help the diagnostic yield of biopsies or even obviate biopsies if it can demonstrate sufficient sensitivity and specificity according the Preservation and Incorporation of Valuable Endoscopic Innovations (PIVI) standards from the American Society for Gastrointestinal Endoscopy. In order to eliminate random mucosal biopsies in BE patients, PIVI recommends that an imaging technology with targeted biopsies should have a per-patient sensitivity of at least 90% and a negative predictive value (NPV) of at least 98% for detecting high-grade dysplasia (HGD) or early esophageal adenocarcinoma (EAC) [12,63].
In the esophagus, Evans et al. showed that OCT could differentiate IMC and HGD from LGD, intermediate-grade dysplasia (IGD) and SIM without dysplasia at the squamocolumnar junction (SCJ) (Evans 2006) [64]. The endoscopic balloon-centering catheter [34] laid the foundation for the commercialization of a GI-focused OCT system (called volumetric laser endomicroscopy, or VLE; NvisionVLE® Imaging System, Ninepoint Medical) [65]. VLE provides real-time volumetric scans with ~3 mm imaging depth and 7 μum axial resolution over a 6 cm stretch of the esophageal inner wall (Figure 3). As will be discussed below, VLE can be used with diagnostic algorithms to detect dysplasia [66] and laser-marking for more precise targeting and co-registry of images with biopsies and better delineation of areas for ablation [67,68]. Similar promising OCT imaging systems are emerging for GI applications including LuminScan ™ (Micro-Tech, Nanjing, CN). On the other hand, imaging probes introduced without balloons may be helpful in real-time imaging during biopsies and may help delineate lateral margins during endoscopic resections [69].
Given the relevance of buried glands and SIM in BE recurrence and WLE’s inability to survey beneath the mucosal surface, OCT can be used as a uniquely-suited tool after radiofrequency ablation (RFA) to evaluate tissue recovery and detect buried or residual glands [12,60,70]. Recently, Tsai et al. demonstrated 3D-OCT’s ability to detect SSIM in BE patients undergoing RFA as a potential predictor for the number of RFA sessions needed for complete eradication of intestinal metaplasia [60,70,71]. They found that 3D-OCT not only provided 80×−100× larger FOV than biopsy, but that its subsurface imaging capabilities revealed that the length of SSIM had a strong correlation with the number of RFA treatments required to achieve complete eradication of intestinal metaplasia [60,71]. VLE has also been shown to differentiate between esophageal squamous cell carcinoma (ESCC) and its precursor, squamous intraepithelial neoplasia (SIN), which are highly prevalent in the non-western world [60,70,71,72]. It was also able to distinguish ESCC limited to the lamina propria and ESCC invading beyond the lamina propria, suggesting that VLE may be able to determine which patients should undergo endoscopic therapy [70].

3.2. Colon

Colorectal cancer (CRC) remains a major cause of death in the US and the second leading cause of cancer death with an estimated 145,600 new cases and an estimated 51,020 deaths for 2019 [71]. The survival rate from GI cancers can be improved by 5- to 10-fold if diagnosed at an early stage [5]. The current recommendation in the US is to screen all individuals at age 45 or earlier if deemed at higher risk for CRC based on certain conditions or family history using specific stool tests, radiological exams or direct visualization by colonoscopy [72]. Colonoscopy allows for the removal of suspect adenomatous lesions but can be time-consuming and has miss rates in the range of 17 to 28% [73]. Thus, advanced imaging modalities could help improve detection. OCT has been demonstrated to differentiate colorectal cancer from adenomatous polyps [74,75], and its ability to reveal the submucosa may allow more nuanced polyp treatments. For example, OCT showed normal submucosa underneath a rectal polyp near the dentate line and altered its management from a full thickness resection to a less invasive endoscopic mucosal resection (EMR) [76]. High-resolution en face colon images from OCT are analogous to the traditional magnified endoscopic view, but do not require contrast agents; they also yield clinically powerful multi-depth visibility [77]. Thus, OCT has the potential to improve polyp detection and management.
Besides neoplastic lesions, 3D-OCT was also able to detect changes in inflammatory conditions of the colon. For example, Crohn’s disease could be distinguished from ulcerative colitis by demonstration of changes in deeper structures not confined to the mucosa, which are hallmarks of Crohn’s Disease [78]. Moreover, OCT was able to demonstrate submucosal fibrosis, with loss of layering, loss of columnar architecture, and superficial edema, which is associated with more severe ulcerative colitis [79].

3.3. Hepatobiliary

Hepatobiliary malignancies account for around 13% of global cancer mortalities and 3% of those in the US, with cholangiocarcinoma (CCA) accounting for 15-20% of hepatobiliary malignancies [80]. The assessment of strictures in the pancreatico-biliary tract can be challenging since they can be benign or cancerous. The current protocol is an endoscopic retrograde cholangiopancreatography (ERCP) and random biopsies or brush cytology [81,82], but the practice often leads to a low diagnostic yield because of the small amount of available tissue and the tortuous anatomy of the bile duct [83,84]. Thus, a low-profile OCT probe was designed to enable passage through the pancreatic duct and demonstrated the ability to differentiate malignant from benign strictures [85,86,87]. Tyberg et al. described that strictures with cholangiocarcinoma showed a hyperreflective surface with loss of inner wall layering, while benign areas had clear delineated ductal wall layering [88]. Joshi, in a 22-patient study, found similar results and was able to quantify the thickness of each duct layer to differentiate benign, inflammatory and malignant tissue types [89]. This ability of OCT to potentially differentiate benign from cancerous lesions may obviate the need to remove tissue in the future.

3.4. Imaging of Microvasculature

In 2014, Tsai et al. demonstrated a 3D OCT angiography (OCT-A) using the biospeckle effect to reveal the subsurface vasculature of BE without exogenous contrast agents [90]. A video introduction of this promising technology in the gastrointestinal tract is available [36]. Revelation of subsurface vasculature could be clinically useful in order to highlight areas to minimize bleeding during an APC hybrid lift ablation (Erbe Elektromedizin, Tübingen, Germany) or peroral endoscopic myotomy, for example. Microvasculature irregularities also appear to be correlated with dysplasia in BE [91,92].
Additionally, OCT-A may help identify specific areas to treat in the rectal wall of chronic radiation proctopathy (CRP) which may develop consequent to pelvic irradiation for various malignancies [93,94]. OCT also may help identify other vascular abnormalities such as Dieulafoy’s lesions, often buried deeply in the submucosal architecture and a source for intermittent but brisk blood loss [95]. Additionally, en face OCT findings have been studied in patients with gastric antral vascular ectasia (GAVE), also a cause of upper GI bleeding, before and after RFA treatment. Lee et al. demonstrated OCT/OCT angiography’s (OCTA’s) ability to differentiate GAVE-associated tissue architecture and microvasculature before and after RFA, namely identifying regions of incomplete ablation hidden under coagulum [96]. Lastly, instantaneous real-time imaging may be used in the future to assess vascular changes associated with food allergies in the gut that are increasingly appreciated as potential causes of diarrhea or irritable bowel syndrome (IBS).

3.5. Small Intestine

The length, tortuous anatomy and folds in the small intestine pose particularly difficult challenges for OCT imaging in the GI tract. Nonetheless, several papers show successful utility of OCT. Masci et al. demonstrated OCT’s ability to differentiate celiac disease based on villous atrophy [97,98], while Kamboj et al. used OCT to reveal the submucosal mass of a duodenal neuroendocrine tumor (NET) [99]. Additionally, Lee et al. used ultrahigh-speed endoscopic OCT to study the structural differences of Crohn’s disease in the terminal ileum (TI), showing that TI with Crohn’s ileitis exhibited irregular mucosal patterns and enlarged villous internal structures [100].

4. New Technologies and Future Applications

4.1. Artificial Intelligence (AI)

From music suggestion algorithms to air traffic management, artificial intelligence (AI) is omnipresent in today’s cutting edge technology. However, unlike programs for checkers and chess, there may be too many predictive parameters for the detection of GI tract neoplasias for a knowledge base to be realistically programmed. Thus, machines must be able to learn. In machine learning, computational methods use experience (i.e. training) to make predictions and improve performance [101]. Deep learning addresses the major pitfall of machine learning, i.e., failure to recognize factors of variance in representation learning, by introducing and solving in terms of other simpler representations [102]. While deep learning can be viewed from a neurally-inspired perspective (often called a “neural network”), having multiple levels of composition need not be neurally inspired [102]. AI has been applied to endoscopic images to help highlight or identify areas of GI cancers [103]. The miss rates for early neoplastic lesions in the GI tract are significant; one retrospective study found a miss rate of 75.2% for gastric superficial neoplasia (GSNs) [104]. Wu et al. demonstrated the feasibility of deep learning in order to detect early cancers and even map the stomach to address the blind spots responsible for many of these misses during gastroscopies [105].
In the colon, AI may be applied to both cancerous and non-cancerous conditions. In ulcerative colitis (UC), a computer-aided diagnosis (CAD) system was able to identify histologic inflammation associated with UC with 74% sensitivity, 97% specificity and 91% accuracy; this is useful, as WLE struggles to differentiate between healing and inflammation (active disease) [106]. For detection of pre-cancerous lesions, polyps may be missed under WLE at a rate of up to 22% [107]. Thus, research has been aimed at training computers to recognize polyp shape [108], texture, and color [109]. The recently described Window Median Depth of valleys Accumulation (WM-DOVA) energy maps system defines polyps as mucosal protrusions and their boundaries in terms of intensity valleys [110,111]. On the basis of this model, which fits better for zenithal views of the polyp, a pixel inside a polyp would be surrounded by high-intensity valleys in the majority of directions, with brighter regions corresponding to pixels with a higher likelihood of a polyp being present [110]. This algorithm worked better for smaller polyps of type 0-II, and the main causes of failure were deviation from the polyp appearance model, low image quality and poor prep [110].
In the esophagus, combining OCT and AI has been aimed to improve detection of neoplasias beyond the present sensitivity of random biopsies per the Seattle Protocol [112]. OCT address this problem with its volumetric rapid imaging, but its interpretation may be time/labor intensive due to the volume of images [113] and may be highly reader-dependent. However, the voluminous library of OCT images could be used for AI training. In BE, the parameters of atypical gland counts, lack of mucosal layering, and surface-to-surface intensity can be used as metrics [66]. Although performed ex-vivo, a study by Leggett et al. used a novel VLE diagnostic algorithm (VLE-DA) for dysplastic detection, producing a sensitivity of 86%, a specificity of 88%, and a diagnostic accuracy of 87% [66], which approaches the PIVI standards. Although the authors used an earlier form of OCT technology’s diagnostic scoring index, they see an in vivo possibility for using an image viewer region of interest (ROI) over a 1 cm longitudinal distance of BE mucosa [66]. Given the importance of subsurface gland-like structures in the esophagus [62], computer-aided detection was devised to highlight these structures to potentially guide biopsies or ablative treatment in the future [114]. Another BE AI paper compared human cross-sectional VLE diagnoses and computer generated en face VLE images and scored the computer’s consistency for layering (kappa=0.62), gland count (kappa=0.62) and subsurface intensity (kappa=0.49) [99].
Intelligent real-time image segmentation (IRIS) (Nine Point Medical, Bedford, MA) is an AI-based image processing software that highlights established VLE features (surface hyper-reflectivity, increased number of glands, and lack of layering) using a color-graded scale superimposed over cross-sectional and en face perspectives (Figure 3) [115].
When VLE-IRIS observed various BE tissues with correlated histology, it demonstrated an ability to recognize increased hyper-reflectivity, number of glands and lack of layering with increasing BE severity [115]; however, the authors note that gradient was not considered for IRIS features, and glands required manual interpretation. Nonetheless, due to VLE-IRIS’s ability to review large amounts of data, further study is required and merited. Beyond structural enhancements, deep learning has yet to be implemented with OCT images in the GI tract as it has, for example, in the eyes [116]. Deep learning has been employed with images in the colon using WLE [117,118] and in the esophagus using WLE with NBI for ESCC [119,120]. Given volumetric OCT advantages over surface imaging of WLE and NBI, deep learning on an OCT platform could be a promising clinical tool. AI has the potential to utilize machine learning to improve neoplastic recognition in the GI tract, where subtleties often lead to high miss rates. Until then, AI has been relegated to a second reader role [103].

4.2. Capsule

Endoscopy requires sedation and can be time and cost demanding. Thus, an office-based delivery system for OCT beyond that of the endoscope is desirable. Early versions of OCT used catheters such as the balloon catheter [32,34,121], but these proved difficult to navigate the flexures of the colon due to their limitation to tubular structures. MEMS scanners [122] and resonant fiber scanning [29,37,123] have also been used, but rotary scanners require proximal motorized pullback, the tethers for which are susceptible to mechanical deformation which degrades scan accuracy along the longitudinal axis [124]. Wireless capsule endoscopy has been used to image the entire GI tract [125], but is not controllable by the OCT operator. However, simply tying a string to the wireless capsule [126,127,128] demonstrated SS-OCT for 3-D circumferential imaging [35,124] tolerable by patients [128]. However, a capsule serving en face OCT would be more desirable, as many markers of disease in the esophagus (mucosal patterns, pit patterns) [129,130] are detectable by such an orientation. A capsule in 2015 was able to combine en face and circumferential luminal scanning to obtain large field coverage (30 cm2) and small field inspection (1 cm2), yielding both volumetric and en face imaging [124]. A more recent capsule paper [131] also incorporated laser marking into tethered capsule endomicroscopy (TCE). Laser marking is a way of co-localizing OCT images with histology [131](120), and the laser marks had a spatial accuracy of better than 0.5 mm. However, small diameter probes sacrifice FOV for transverse resolution, which is a potential drawback of en face capsule technology [124]. In 2017, a study involving Harvard Medical School and MIT demonstrated a cycloid scanner with a tethered capsule for ultrahigh speed side-viewing OCT in vivo and also produced volumetric images of vasculature [132].

5. Conclusion/Outlook for OCT

OCT is a rapid, volumetric, nearly microscopically resolved, subsurface endoscopic imaging modality that addresses many of the pitfalls associated with WLE and NBI. OCT has been rigorously studied in the esophagus due to its tubular anatomy and is starting to gain traction throughout the rest of the GI tract including the stomach, small intestine and colon. OCT is compatible with a fleet of accessories including laser markers, autofluorescence techniques and the traditional acoustic ultrasound. Additionally, OCT has the potential to be performed using an office-based capsule, negating the need for the burdens of sedation. Given the copious amounts of data produced by OCT, it is also a promising platform for the implementation of machine learning to achieve a level of autodetection akin to or even surpassing that of the skilled endoscopist. While machine learning in OCT still has a long way to go before it reaches such a skillful level, it can nonetheless play the role of second reader. The plethora of benefits from OCT, including accessorization, ease of delivery and machine learning, uniquely position this technology to address the disease burden of gastrointestinal maladies.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Seward, E.; Lumley, S. Endoscopy provision: Meeting the challenges. Frontline Gastroenterol. 2017, 8, 90. [Google Scholar] [CrossRef] [PubMed]
  2. Kurniawan, N.; Keuchel, M. Flexible Gastro-intestinal Endoscopy—Clinical Challenges and Technical Achievements. Comput. Struct. Biotechnol. J. 2017, 15, 168–179. [Google Scholar] [CrossRef] [PubMed]
  3. Bond, A.; Sarkar, S. New technologies and techniques to improve adenoma detection in colonoscopy. World J. Gastrointest. Endosc. 2015, 7, 969–980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Groff, R.J.; Nash, R.; Ahnen, D.J. Significance of serrated polyps of the colon. Curr. Gastroenterol. Rep. 2008, 10, 490–498. [Google Scholar] [CrossRef] [PubMed]
  5. A.C. Cancer Facts and Figures—2017. Available online: https://www.cancer.org/research/cancer-facts-statistics/all-cancer-facts-figures/cancer-facts-figures-2017.html. (accessed on 17 June 2019).
  6. Song, L.M.W.K.; Adler, D.G.; Chand, B.; Conway, J.D.; Croffie, J.M.B.; DiSario, J.A.; Mishkin, D.S.; Shah, R.J.; Somogyi, L.; Tierney, W.M. Chromoendoscopy. Gastrointest. Endosc. 2007, 66, 639–649. [Google Scholar] [CrossRef] [PubMed]
  7. Manfredi, M.A.; Dayyeh, B.K.A.; Bhat, Y.M.; Chauhan, S.S.; Gottlieb, K.T.; Hwang, J.H.; Komanduri, S.; Konda, V.; Lo, S.K.; Maple, J.T. Electronic chromoendoscopy. Gastrointest. Endosc. 2015, 81, 249–261. [Google Scholar] [CrossRef] [PubMed]
  8. Kiesslich, R.; Jung, M. Magnification endoscopy: Does it improve mucosal surface analysis for the diagnosis of gastrointestinal neoplasias? Endoscopy 2002, 34, 819–822. [Google Scholar] [CrossRef]
  9. Song, L.-M.W.K.; Banerjee, S.; Desilets, D.; Diehl, D.L.; Farraye, F.A.; Kaul, V.; Kethu, S.R.; Kwon, R.S.; Mamula, P.; Pedrosa, M.C. Autofluorescence imaging. Gastrointest. Endosc. 2011, 73, 647–650. [Google Scholar] [CrossRef]
  10. Zuccaro, G.; Gladkova, N.; Vargo, J.; Feldchtein, F.; Zagaynova, E.; Conwell, D.; Falk, G.; Goldblum, J.; Dumot, J.; Ponsky, J. Optical coherence tomography of the esophagus and proximal stomach in health and disease. Am. J. Gastroenterol. 2001, 96, 2633. [Google Scholar] [CrossRef]
  11. Chauhan, S.S.; Dayyeh, B.K.A.; Bhat, Y.M.; Gottlieb, K.T.; Hwang, J.H.; Komanduri, S.; Konda, V.; Lo, S.K.; Manfredi, M.A.; Maple, J.T. Confocal laser endomicroscopy. Gastrointest. Endosc. 2014, 80, 928–938. [Google Scholar] [CrossRef]
  12. Tsai, T.H.; Leggett, C.L.; Trindade, A.J.; Sethi, A.; Swager, A.F.; Joshi, V.; Bergman, J.J.; Mashimo, H.; Nishioka, N.S.; Namati, E. Optical coherence tomography in gastroenterology: A review and future outlook. J. Biomed. Opt. 2017, 22, 1–17. [Google Scholar] [CrossRef] [PubMed]
  13. Takada, K.; Yokohama, I.; Chida, K.; Noda, J. New measurement system for fault location in optical waveguide devices based on an interferometric technique. Appl. Opt. 1987, 26, 1603–1606. [Google Scholar] [CrossRef] [PubMed]
  14. Gilgen, H.H.; Novak, R.P.; Salathe, R.P.; Hodel, W.; Beaud, P. Submillimeter optical reflectometry. J. Lightw. Technol. 1989, 7, 1225–1233. [Google Scholar] [CrossRef]
  15. Huang, D.; Swanson, E.A.; Lin, C.P.; Schuman, J.S.; Stinson, W.G.; Chang, W.; Hee, M.R.; Flotte, T.; Gregory, K.; Puliafito, C.A.; et al. Optical coherence tomography. Science 1991, 254, 1178–1181. [Google Scholar] [CrossRef] [PubMed]
  16. Tearney, G.J.; Brezinski, M.E.; Bouma, B.E.; Boppart, S.A.; Pitris, C.; Southern, J.F.; Fujimoto, J.G. In vivo endoscopic optical biopsy with optical coherence tomography. Science 1997, 276, 2037–2039. [Google Scholar] [CrossRef] [PubMed]
  17. Swanson, E.A.; Izatt, J.A.; Hee, M.R.; Huang, D.; Lin, C.P.; Schuman, J.S.; Puliafito, C.A.; Fujimoto, J.G. In vivo retinal imaging by optical coherence tomography. Opt. Lett. 1993, 18, 1864–1866. [Google Scholar] [CrossRef] [PubMed]
  18. Fercher, A.F.; Hitzenberger, C.K.; Kamp, G.; El-Zaiat, S.Y. Measurement of intraocular distances by backscattering spectral interferometry. Opt. Commun. 1995, 117, 43–48. [Google Scholar] [CrossRef]
  19. Cense, B.; Nassif, N.A.; Chen, T.C.; Pierce, M.C.; Yun, S.-H.; Park, B.H.; Bouma, B.E.; Tearney, G.J.; de Boer, J.F. Ultrahigh-resolution high-speed retinal imaging using spectral-domain optical coherence tomography. Opt. Express 2004, 12, 2435–2447. [Google Scholar] [CrossRef]
  20. Wojtkowski, M.; Bajraszewski, T.; Gorczyńska, I.; Targowski, P.; Kowalczyk, A.; Wasilewski, W.; Radzewicz, C. Ophthalmic imaging by spectral optical coherence tomography. Am. J. Ophthalmol. 2004, 138, 412–419. [Google Scholar] [CrossRef]
  21. Popescu, D.P.; Choo-Smith, L.-P.I.; Flueraru, C.; Mao, Y.; Chang, S.; Disano, J.; Sherif, S.; Sowa, M.G. Optical coherence tomography: Fundamental principles, instrumental designs and biomedical applications. Biophys. Rev. 2011, 3, 155. [Google Scholar] [CrossRef]
  22. Swanson, E.A.; Huang, D.; Hee, M.R.; Fujimoto, J.G.; Lin, C.P.; Puliafito, C.A. High-speed optical coherence domain reflectometry. Opt. Lett. 1992, 17, 151–153. [Google Scholar] [CrossRef] [PubMed]
  23. Tsai, T.-H.; Fujimoto, G.J.; Mashimo, H. Endoscopic Optical Coherence Tomography for Clinical Gastroenterology. Diagnostics 2014, 4, 57–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Choma, M.; Sarunic, M.; Yang, C.; Izatt, J. Sensitivity advantage of swept source and Fourier domain optical coherence tomography. Opt. Express 2003, 11, 2183–2189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. de Boer, J.F.; Cense, B.; Park, B.H.; Pierce, M.C.; Tearney, G.J.; Bouma, B.E. Improved signal-to-noise ratio in spectral-domain compared with time-domain optical coherence tomography. Opt. Lett. 2003, 28, 2067–2069. [Google Scholar] [CrossRef] [PubMed]
  26. Leitgeb, R.; Hitzenberger, C.K.; Fercher, A.F. Performance of fourier domain vs. time domain optical coherence tomography. Opt. Express 2003, 11, 889–894. [Google Scholar] [CrossRef]
  27. Ko, T.H.; Adler, D.C.; Fujimoto, J.G.; Mamedov, D.; Prokhorov, V.; Shidlovski, V.; Yakubovich, S. Ultrahigh resolution optical coherence tomography imaging with a broadband superluminescent diode light source. Opt. Express 2004, 12, 2112–2119. [Google Scholar] [CrossRef] [PubMed]
  28. Robles, F.E.; Chowdhury, S.; Wax, A. Assessing hemoglobin concentration using spectroscopic optical coherence tomography for feasibility of tissue diagnostics. Biomed. Opt. Express 2010, 1, 310–317. [Google Scholar] [CrossRef] [Green Version]
  29. Tsai, T.H.; Potsaid, B.; Kraus, M.F.; Zhou, C.; Tao, Y.K.; Hornegger, J.; Fujimoto, J.G. Piezoelectric-transducer-based miniature catheter for ultrahigh-speed endoscopic optical coherence tomography. Biomed. Opt. Express 2011, 2, 2438–2448. [Google Scholar] [CrossRef]
  30. Wieser, W.; Biedermann, B.R.; Klein, T.; Eigenwillig, C.M.; Huber, R. Multi-Megahertz OCT: High quality 3D imaging at 20 million A-scans and 4.5 GVoxels per second. Opt. Express 2010, 18, 14685–14704. [Google Scholar] [CrossRef]
  31. Wang, T.; Pfeiffer, T.; Regar, E.; Wieser, W.; van Beusekom, H.; Lancee, C.T.; Springeling, G.; Krabbendam-Peters, I.; van der Steen, A.F.W.; Huber, R.; et al. Heartbeat OCT and Motion-Free 3D In Vivo Coronary Artery Microscopy. JACC Cardiovasc. Imaging 2016, 9, 622–623. [Google Scholar] [CrossRef]
  32. Vakoc, B.J.; Shishko, M.; Yun, S.H.; Oh, W.Y.; Suter, M.J.; Desjardins, A.E.; Evans, J.A.; Nishioka, N.S.; Tearney, G.J.; Bouma, B.E. Comprehensive esophageal microscopy by using optical frequency-domain imaging (with video). Gastrointest. Endosc. 2007, 65, 898–905. [Google Scholar] [CrossRef] [PubMed]
  33. Adler, D.C.; Chen, Y.; Huber, R.; Schmitt, J.; Connolly, J.; Fujimoto, J.G. Three-dimensional endomicroscopy using optical coherence tomography. Nat. Photonics 2007, 1, 709. [Google Scholar] [CrossRef]
  34. Suter, M.J.; Vakoc, B.J.; Yachimski, P.S.; Shishkov, M.; Lauwers, G.Y.; Mino-Kenudson, M.; Bouma, B.E.; Nishioka, N.S.; Tearney, G.J. Comprehensive microscopy of the esophagus in human patients with optical frequency domain imaging. Gastrointest. Endosc. 2008, 68, 745–753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Gora, M.J.; Sauk, J.S.; Carruth, R.W.; Gallagher, K.A.; Suter, M.J.; Nishioka, N.S.; Kava, L.E.; Rosenberg, M.; Bouma, B.E.; Tearney, G.J. Tethered capsule endomicroscopy enables less invasive imaging of gastrointestinal tract microstructure. Nat. Med. 2013, 19, 238–240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Supplementary Material. Available online: https://www.gastrojournal.org/article/S0016-5085(14)01073-7/fulltext#appsec1 (accessed on 19 July 2019).
  37. Zhang, N.; Tsai, T.-H.; Ahsen, O.O.; Liang, K.; Lee, H.-C.; Xue, P.; Li, X.; Fujimoto, J.G. Compact piezoelectric transducer fiber scanning probe for optical coherence tomography. Opt. Lett. 2014, 39, 186–188. [Google Scholar] [CrossRef] [PubMed]
  38. Gora, M.J.; Suter, M.J.; Tearney, G.J.; Li, X. Endoscopic optical coherence tomography: Technologies and clinical applications [Invited]. Biomed. Opt. Express 2017, 8, 2405–2444. [Google Scholar] [CrossRef] [PubMed]
  39. Cogliati, A.; Canavesi, C.; Hayes, A.; Tankam, P.; Duma, V.-F.; Santhanam, A.; Thompson, K.P.; Rolland, J.P. MEMS-based handheld scanning probe with pre-shaped input signals for distortion-free images in Gabor-domain optical coherence microscopy. Opt. Express 2016, 24, 13365–13374. [Google Scholar] [CrossRef] [PubMed]
  40. Mu, X.; Zhou, G.; Yu, H.; Du, Y.; Feng, H.; Tsai, J.M.; Chau, F.S. Compact MEMS-driven pyramidal polygon reflector for circumferential scanned endoscopic imaging probe. Opt. Express 2012, 20, 6325–6339. [Google Scholar] [CrossRef] [PubMed]
  41. Wu, J.; Conry, M.; Gu, C.; Wang, F.; Yaqoob, Z.; Yang, C. Paired-angle-rotation scanning optical coherence tomography forward-imaging probe. Opt. Lett. 2006, 31, 1265–1267. [Google Scholar] [CrossRef]
  42. Ahsen, O.O.; Liang, K.; Lee, H.C.; Giacomelli, M.G.; Wang, Z.; Potsaid, B.; Figueiredo, M.; Huang, Q.; Jayaraman, V.; Fujimoto, J.G.; et al. Assessment of Barrett’s esophagus and dysplasia with ultrahigh-speed volumetric en face and cross-sectional optical coherence tomography. Endoscopy 2019, 51, 355–359. [Google Scholar] [CrossRef]
  43. Pan, Y.; Xie, H.; Fedder, G.K. Endoscopic optical coherence tomography based on a microelectromechanical mirror. Opt. Lett. 2001, 26, 1966–1968. [Google Scholar] [CrossRef] [PubMed]
  44. Seibel, E.J.; Smithwick, Q.Y. Unique features of optical scanning, single fiber endoscopy. Lasers Surg. Med. 2002, 30, 177–183. [Google Scholar] [CrossRef] [PubMed]
  45. Moon, S.; Lee, S.-W.; Rubinstein, M.; Wong, B.J.F.; Chen, Z. Semi-resonant operation of a fiber-cantilever piezotube scanner for stable optical coherence tomography endoscope imaging. Opt. Express 2010, 18, 21183–21197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Ryu, S.Y.; Choi, H.Y.; Na, J.; Choi, E.S.; Lee, B.H. Combined system of optical coherence tomography and fluorescence spectroscopy based on double-cladding fiber. Opt. Lett. 2008, 33, 2347–2349. [Google Scholar] [CrossRef] [PubMed]
  47. Tumlinson, A.R.; Hariri, L.P.; Utzinger, U.; Barton, J.K. Miniature endoscope for simultaneous optical coherence tomography and laser-induced fluorescence measurement. Appl. Opt. 2004, 43, 113–121. [Google Scholar] [CrossRef] [Green Version]
  48. Li, Y.; Jing, J.; Yu, J.; Zhang, B.; Huo, T.; Yang, Q.; Chen, Z. Multimodality endoscopic optical coherence tomography and fluorescence imaging technology for visualization of layered architecture and subsurface microvasculature. Opt. Lett. 2018, 43, 2074–2077. [Google Scholar] [CrossRef]
  49. Li, X.; Yin, J.; Hu, C.; Zhou, Q.; Shung, K.K.; Chen, Z. High-resolution coregistered intravascular imaging with integrated ultrasound and optical coherence tomography probe. Appl. Phys. Lett. 2010, 97, 133702. [Google Scholar] [CrossRef] [Green Version]
  50. Yang, Y.; Li, X.; Wang, T.; Kumavor, P.D.; Aguirre, A.; Shung, K.K.; Zhou, Q.; Sanders, M.; Brewer, M.; Zhu, Q. Integrated optical coherence tomography, ultrasound and photoacoustic imaging for ovarian tissue characterization. Biomed. Opt. Express 2011, 2, 2551–2561. [Google Scholar] [CrossRef] [Green Version]
  51. de Boer, J.F.; Hitzenberger, C.K.; Yasuno, Y. Polarization sensitive optical coherence tomography—A review [Invited]. Biomed. Opt. Express 2017, 8, 1838–1873. [Google Scholar] [CrossRef]
  52. Antonelli, M.R.; Pierangelo, A.; Novikova, T.; Validire, P.; Benali, A.; Gayet, B.; De Martino, A. Mueller matrix imaging of human colon tissue for cancer diagnostics: How Monte Carlo modeling can help in the interpretation of experimental data. Opt. Express 2010, 18, 10200–10208. [Google Scholar] [CrossRef]
  53. Steele, D.; Baig, K.K.K.; Peter, S. Evolving screening and surveillance techniques for Barrett’s esophagus. World J. Gastroenterol. 2019, 25, 2045–2057. [Google Scholar] [CrossRef] [PubMed]
  54. Maes, S.; Sharma, P.; Bisschops, R. Review: Surveillance of patients with Barrett oesophagus. Best Pract. Res. Clin. Gastroenterol. 2016, 30, 901–912. [Google Scholar] [CrossRef] [PubMed]
  55. Hur, C.; Miller, M.; Kong, C.Y.; Dowling, E.C.; Nattinger, K.J.; Dunn, M.; Feuer, E.J. Trends in esophageal adenocarcinoma incidence and mortality. Cancer 2013, 119, 1149–1158. [Google Scholar] [CrossRef] [PubMed]
  56. Lagergren, J.; Lagergren, P. Recent developments in esophageal adenocarcinoma. CA Cancer J. Clin. 2013, 63, 232–248. [Google Scholar] [CrossRef] [PubMed]
  57. Souza, R.F.; Morales, C.P.; Spechler, S.J. Review article: A conceptual approach to understanding the molecular mechanisms of cancer development in Barrett’s oesophagus. Aliment. Pharmacol. Ther. 2001, 15, 1087–1100. [Google Scholar] [CrossRef]
  58. Weston, A.P.; Banerjee, S.K.; Sharma, P.; Tran, T.M.; Richards, R.; Cherian, R. p53 protein overexpression in low grade dysplasia (LGD) in Barrett’s esophagus: Immunohistochemical marker predictive of progression. Am. J. Gastroenterol. 2001, 96, 1355–1362. [Google Scholar] [CrossRef]
  59. Shaheen, N.J.; Falk, G.W.; Iyer, P.G.; Gerson, L.B. ACG Clinical Guideline: Diagnosis and Management of Barrett’s Esophagus. Am. J. Gastroenterol. 2016, 111, 30. [Google Scholar] [CrossRef]
  60. Zhou, C.; Tsai, T.H.; Lee, H.C.; Kirtane, T.; Figueiredo, M.; Tao, Y.K.; Ahsen, O.O.; Adler, D.C.; Schmitt, J.M.; Huang, Q.; et al. Characterization of buried glands before and after radiofrequency ablation by using 3-dimensional optical coherence tomography (with videos). Gastrointest. Endosc. 2012, 76, 32–40. [Google Scholar] [CrossRef] [Green Version]
  61. Pouw, R.E.; Gondrie, J.J.; Rygiel, A.M.; Sondermeijer, C.M.; ten Kate, F.J.; Odze, R.D.; Vieth, M.; Krishnadath, K.K.; Bergman, J.J. Properties of the neosquamous epithelium after radiofrequency ablation of Barrett’s esophagus containing neoplasia. Am. J. Gastroenterol. 2009, 104, 1366–1373. [Google Scholar] [CrossRef]
  62. Mashimo, H. Subsquamous intestinal metaplasia after ablation of Barrett’s esophagus: Frequency and importance. Curr. Opin. Gastroenterol. 2013, 29, 454–459. [Google Scholar] [CrossRef]
  63. Sharma, P.; Savides, T.J.; Canto, M.I.; Corley, D.A.; Falk, G.W.; Goldblum, J.R.; Wang, K.K.; Wallace, M.B.; Wolfsen, H.C. The American Society for Gastrointestinal Endoscopy PIVI (Preservation and Incorporation of Valuable Endoscopic Innovations) on imaging in Barrett’s Esophagus. Gastrointest. Endosc. 2012, 76, 252–254. [Google Scholar] [CrossRef] [PubMed]
  64. Evans, J.A.; Poneros, J.M.; Bouma, B.E.; Bressner, J.; Halpern, E.F.; Shishkov, M.; Lauwers, G.Y.; Mino-Kenudson, M.; Nishioka, N.S.; Tearney, G.J. Optical coherence tomography to identify intramucosal carcinoma and high-grade dysplasia in Barrett’s esophagus. Clin. Gastroenterol. Hepatol. 2006, 4, 38–43. [Google Scholar] [CrossRef]
  65. Trindade, A.J.; Vamadevan, A.S.; Sejpal, D.V. Finding a needle in a haystack: Use of volumetric laser endomicroscopy in targeting focal dysplasia in long-segment Barrett’s esophagus. Gastrointest. Endosc. 2015, 82, 756–757. [Google Scholar] [CrossRef] [PubMed]
  66. Leggett, C.L.; Gorospe, E.C.; Chan, D.K.; Muppa, P.; Owens, V.; Smyrk, T.C.; Anderson, M.; Lutzke, L.S.; Tearney, G.; Wang, K.K. Comparative diagnostic performance of volumetric laser endomicroscopy and confocal laser endomicroscopy in the detection of dysplasia associated with Barrett’s esophagus. Gastrointest. Endosc. 2016, 83, 880–888. [Google Scholar] [CrossRef]
  67. Swager, A.F.; Tearney, G.J.; Leggett, C.L.; van Oijen, M.G.H.; Meijer, S.L.; Weusten, B.L.; Curvers, W.L.; Bergman, J. Identification of volumetric laser endomicroscopy features predictive for early neoplasia in Barrett’s esophagus using high-quality histological correlation. Gastrointest. Endosc. 2017, 85, 918–926. [Google Scholar] [CrossRef] [PubMed]
  68. Suter, M.J.; Jillella, P.A.; Vakoc, B.J.; Halpern, E.F.; Mino-Kenudson, M.; Lauwers, G.Y.; Bouma, B.E.; Nishioka, N.S.; Tearney, G.J. Image-guided biopsy in the esophagus through comprehensive optical frequency domain imaging and laser marking: A study in living swine. Gastrointest. Endosc. 2010, 71, 346–353. [Google Scholar] [CrossRef] [PubMed]
  69. Ahsen, O.O.; Lee, H.C.; Liang, K.; Wang, Z.; Figueiredo, M.; Huang, Q.; Potsaid, B.; Jayaraman, V.; Fujimoto, J.G.; Mashimo, H. Ultrahigh-speed endoscopic optical coherence tomography and angiography enables delineation of lateral margins of endoscopic mucosal resection: A case report. Ther. Adv. Gastroenterol. 2017, 10, 931–936. [Google Scholar] [CrossRef]
  70. Trindade, A.J.; Benias, P.C.; Inamdar, S.; Fan, C.; Sethi, A.; Fukami, N.; Kahn, A.; Kahaleh, M.; Andalib, I.; Sejpal, D.V.; et al. Use of volumetric laser endomicroscopy for determining candidates for endoscopic therapy in superficial esophageal squamous cell carcinoma. United Eur. Gastroenterol. J. 2018, 6, 838–845. [Google Scholar] [CrossRef]
  71. Colorectal Cancer Facts and Figures 2017–2019. Available online: https://www.cancer.org/content/dam/cancer-org/research/cancer-facts-and-statistics/colorectal-cancer-facts-and-figures/colorectal-cancer-facts-and-figures-2017-2019.pdf (accessed on 17 June 2019).
  72. Wolf, A.M.D.; Fontham, E.T.H.; Church, T.R.; Flowers, C.R.; Guerra, C.E.; LaMonte, S.J.; Etzioni, R.; McKenna, M.T.; Oeffinger, K.C.; Shih, Y.-C.T.; et al. Colorectal cancer screening for average-risk adults: 2018 guideline update from the American Cancer Society. CA Cancer J. Clin. 2018, 68, 250–281. [Google Scholar] [CrossRef]
  73. Kim, N.H.; Jung, Y.S.; Jeong, W.S.; Yang, H.-J.; Park, S.-K.; Choi, K.; Park, D.I. Miss rate of colorectal neoplastic polyps and risk factors for missed polyps in consecutive colonoscopies. Intest. Res. 2017, 15, 411–418. [Google Scholar] [CrossRef]
  74. Pfau, P.R.; Sivak, M.V., Jr.; Chak, A.; Kinnard, M.; Wong, R.C.; Isenberg, G.A.; Izatt, J.A.; Rollins, A.; Westphal, V. Criteria for the diagnosis of dysplasia by endoscopic optical coherence tomography. Gastrointest. Endosc. 2003, 58, 196–202. [Google Scholar] [CrossRef] [PubMed]
  75. Zagaynova, E.; Gladkova, N.; Shakhova, N.; Gelikonov, G.; Gelikonov, V. Endoscopic OCT with forward-looking probe: Clinical studies in urology and gastroenterology. J. Biophotonics 2008, 1, 114–128. [Google Scholar] [CrossRef] [PubMed]
  76. Trindade, A.J.; Sultan, K.; Vamadevan, A.S.; Fan, C.; Sejpal, D.V. Successful use of volumetric laser endomicroscopy in imaging a rectal polyp. Ther. Adv. Gastroenterol. 2015, 9, 128–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Liang, K.; Ahsen, O.O.; Wang, Z.; Lee, H.-C.; Liang, W.; Potsaid, B.M.; Tsai, T.-H.; Giacomelli, M.G.; Jayaraman, V.; Mashimo, H.; et al. Endoscopic forward-viewing optical coherence tomography and angiography with MHz swept source. Opt. Lett. 2017, 42, 3193–3196. [Google Scholar] [CrossRef] [PubMed]
  78. Shen, B.; Zuccaro, G., Jr.; Gramlich, T.L.; Gladkova, N.; Trolli, P.; Kareta, M.; Delaney, C.P.; Connor, J.T.; Lashner, B.A.; Bevins, C.L.; et al. In vivo colonoscopic optical coherence tomography for transmural inflammation in inflammatory bowel disease. Clin. Gastroenterol. Hepatol. 2004, 2, 1080–1087. [Google Scholar] [CrossRef]
  79. Adler, D.C.; Zhou, C.; Tsai, T.-H.; Schmitt, J.; Huang, Q.; Mashimo, H.; Fujimoto, J.G. Three-dimensional endomicroscopy of the human colon using optical coherence tomography. Opt. Express 2009, 17, 784–796. [Google Scholar] [CrossRef] [PubMed]
  80. Blechacz, B. Cholangiocarcinoma: Current Knowledge and New Developments. Gut Liver 2017, 11, 13–26. [Google Scholar] [CrossRef]
  81. Pugliese, V.; Pujic, N.; Saccomanno, S.; Gatteschi, B.; Pera, C.; Aste, H.; Ferrara, G.B.; Nicolò, G. Pancreatic intraductal sampling during ERCP in patients with chronic pancreatitis and pancreatic cancer: Cytologic studies and k-ras-2 codon 12 molecular analysis in 47 cases. Gastrointest. Endosc. 2001, 54, 595–599. [Google Scholar] [CrossRef]
  82. Rösch, T.; Hofrichter, K.; Frimberger, E.; Meining, A.; Born, P.; Weigert, N.; Allescher, H.-D.; Classen, M.; Barbur, M.; Schenck, U.; et al. ERCP or EUS for tissue diagnosis of biliary strictures? A prospective comparative study. Gastrointest. Endosc. 2004, 60, 390–396. [Google Scholar] [CrossRef]
  83. Selvaggi, S.M. Biliary brushing cytology. Cytopathology 2004, 15, 74–79. [Google Scholar] [CrossRef]
  84. de Bellis, M.; Sherman, S.; Fogel, E.L.; Cramer, H.; Chappo, J.; McHenry, L.; Watkins, J.L.; Lehman, G.A. Tissue sampling at ERCP in suspected malignant biliary strictures (Part 2). Gastrointest. Endosc. 2002, 56, 720–730. [Google Scholar] [CrossRef]
  85. Seitz, U.; Freund, J.; Jaeckle, S.; Feldchtein, F.; Bohnacker, S.; Thonke, F.; Gladkova, N.; Brand, B.; Schroder, S.; Soehendra, N. First in vivo optical coherence tomography in the human bile duct. Endoscopy 2001, 33, 1018–1021. [Google Scholar] [CrossRef] [PubMed]
  86. Poneros, J.M.; Tearney, G.J.; Shiskov, M.; Kelsey, P.B.; Lauwers, G.Y.; Nishioka, N.S.; Bouma, B.E. Optical coherence tomography of the biliary tree during ERCP. Gastrointest. Endosc. 2002, 55, 84–88. [Google Scholar] [CrossRef]
  87. Testoni, P.A.; Mariani, A.; Mangiavillano, B.; Arcidiacono, P.G.; Di Pietro, S.; Masci, E. Intraductal optical coherence tomography for investigating main pancreatic duct strictures. Am. J. Gastroenterol. 2007, 102, 269–274. [Google Scholar] [CrossRef] [PubMed]
  88. Tyberg, A.; Xu, M.-M.; Gaidhane, M.; Sharaiha, R.Z.; Kahaleh, M. Tu2004—Second Generation Optical Coherence Tomography: Preliminary Experience in Pancreatic and Biliary Strictures. Gastroenterology 2017, 152, S1032–S1033. [Google Scholar] [CrossRef]
  89. Joshi, V.; Patel, S.N.; Vanderveldt, H.; Oliva, I.; Raijman, I.; Molina, C.; Carr-Locke, D.L. Mo1963 A Pilot Study of Safety and Efficacy of Directed Cannulation with a Low Profile Catheter (LP) and Imaging Characteristics of Bile Duct Wall Using Optical Coherance Tomography (OCT) for Indeterminate Biliary Strictures Initial Report on In-Vivo Evaluation During ERCP. Gastrointest. Endosc. 2017, 85, AB496–AB497. [Google Scholar] [CrossRef]
  90. Tsai, T.-H.; Ahsen, O.O.; Lee, H.-C.; Liang, K.; Figueiredo, M.; Tao, Y.K.; Giacomelli, M.G.; Potsaid, B.M.; Jayaraman, V.; Huang, Q. Endoscopic optical coherence angiography enables 3-dimensional visualization of subsurface microvasculature. Gastroenterology 2014, 147, 1219–1221. [Google Scholar] [CrossRef]
  91. Lee, M.M.; Enns, R. Narrow band imaging in gastroesophageal reflux disease and Barrett’s esophagus. Can. J. Gastroenterol. Hepatol. 2009, 23, 84–87. [Google Scholar] [CrossRef]
  92. Singh, R.; Anagnostopoulos, G.K.; Yao, K.; Karageorgiou, H.; Fortun, P.J.; Shonde, A.; Garsed, K.; Kaye, P.V.; Hawkey, C.J.; Ragunath, K. Narrow-band imaging with magnification in Barrett’s esophagus: Validation of a simplified grading system of mucosal morphology patterns against histology. Endoscopy 2008, 40, 457–463. [Google Scholar] [CrossRef]
  93. Zhou, C.; Adler, D.C.; Becker, L.; Chen, Y.; Tsai, T.-H.; Figueiredo, M.; Schmitt, J.M.; Fujimoto, J.G.; Mashimo, H. Effective treatment of chronic radiation proctitis using radiofrequency ablation. Ther. Adv. Gastroenterol. 2009, 2, 149–156. [Google Scholar] [CrossRef]
  94. Lenz, L.; Rohr, R.; Nakao, F.; Libera, E.; Ferrari, A. Chronic radiation proctopathy: A practical review of endoscopic treatment. World J. Gastrointest. Surg. 2016, 8, 151–160. [Google Scholar] [CrossRef] [PubMed]
  95. Nojkov, B.; Cappell, M.S. Gastrointestinal bleeding from Dieulafoy’s lesion: Clinical presentation, endoscopic findings, and endoscopic therapy. World J. Gastrointest. Endosc. 2015, 7, 295–307. [Google Scholar] [CrossRef] [PubMed]
  96. Lee, H.-C.; Ahsen, O.O.; Liang, K.; Wang, Z.; Figueiredo, M.; Potsaid, B.; Jayaraman, V.; Huang, Q.; Mashimo, H.; Fujimoto, J.G. Sa2034 Novel Ultrahigh Speed Endoscopic OCT Angiography Identifies Both Architectural and Microvascular Changes in Patients with Gastric Antral Vascular Ectasia (GAVE) Undergoing Radiofrequency Ablation (RFA) Treatment. Gastroenterology 2016, 150, S435–S436. [Google Scholar] [CrossRef]
  97. Masci, E.; Mangiavillano, B.; Albarello, L.; Mariani, A.; Doglioni, C.; Testoni, P.A. Pilot study on the correlation of optical coherence tomography with histology in celiac disease and normal subjects. J. Gastroenterol. Hepatol. 2007, 22, 2256–2260. [Google Scholar] [CrossRef] [PubMed]
  98. Masci, E.; Mangiavillano, B.; Barera, G.; Parma, B.; Albarello, L.; Mariani, A.; Doglioni, C.; Testoni, P.A. Optical coherence tomography in pediatric patients: A feasible technique for diagnosing celiac disease in children with villous atrophy. Dig. Liver Dis. 2009, 41, 639–643. [Google Scholar] [CrossRef]
  99. Kamboj, A.K.; Chan, D.K.; Zakko, L.; Visrodia, K.; Otaki, F.; Lutzke, L.S.; Wang, K.K.; Leggett, C.L. Mo2006 Detection of Barrett’s Esophagus Dysplasia Using a Novel Volumetric Laser Endomicroscopy Computer Algorithm. Gastrointest. Endosc. 2017, 85, AB518. [Google Scholar] [CrossRef]
  100. Lee, H.-C.; Ahsen, O.O.; Liang, K.; Wang, Z.; Figueiredo, M.; Giacomelli, M.G.; Potsaid, B.; Huang, Q.; Mashimo, H.; Fujimoto, J.G. Endoscopic optical coherence tomography angiography microvascular features associated with dysplasia in Barrett’s esophagus (with video). Gastrointest. Endosc. 2017, 86, 476–484. [Google Scholar] [CrossRef]
  101. Mohri, M.; Rostamizadeh, A.; Talwalkar, A. Foundations of Machine Learning; MIT Press: London, UK, 2018. [Google Scholar]
  102. Goodfellow, I.; Bengio, Y.; Courville, A. Deep Learning; MIT Press: London, UK, 2016. [Google Scholar]
  103. Byrne, M.F. Artificial intelligence and the future of endoscopy: Should we be quietly excited? Endoscopy 2019, 51, 511–512. [Google Scholar] [CrossRef]
  104. Shimodate, Y.; Mizuno, M.; Doi, A.; Nishimura, N.; Mouri, H.; Matsueda, K.; Yamamoto, H. Gastric superficial neoplasia: High miss rate but slow progression. Endosc. Int. Open 2017, 5, E722–E726. [Google Scholar] [CrossRef]
  105. Wu, L.; Zhou, W.; Wan, X.; Zhang, J.; Shen, L.; Hu, S.; Ding, Q.; Mu, G.; Yin, A.; Huang, X.; et al. A deep neural network improves endoscopic detection of early gastric cancer without blind spots. Endoscopy 2019, 51, 522–531. [Google Scholar] [CrossRef] [Green Version]
  106. Maeda, Y.; Kudo, S.-E.; Mori, Y.; Misawa, M.; Ogata, N.; Sasanuma, S.; Wakamura, K.; Oda, M.; Mori, K.; Ohtsuka, K. Fully automated diagnostic system with artificial intelligence using endocytoscopy to identify the presence of histologic inflammation associated with ulcerative colitis (with video). Gastrointest. Endosc. 2019, 89, 408–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. van Rijn, J.C.; Reitsma, J.B.; Stoker, J.; Bossuyt, P.M.; van Deventer, S.J.; Dekker, E. Polyp miss rate determined by tandem colonoscopy: A systematic review. Am. J. Gastroenterol. 2006, 101, 343–350. [Google Scholar] [CrossRef] [PubMed]
  108. Tajbakhsh, N.; Gurudu, S.R.; Liang, J. Automatic polyp detection using global geometric constraints and local intensity variation patterns. In International Conference on Medical Image Computing and Computer-Assisted Intervention; Springer: Cham, Switzerland, 2014; Volume 17, pp. 179–187. [Google Scholar]
  109. Bernal, J.; Sánchez, J.; Vilarino, F. Towards automatic polyp detection with a polyp appearance model. Pattern Recognit. 2012, 45, 3166–3182. [Google Scholar] [CrossRef]
  110. Fernandez-Esparrach, G.; Bernal, J.; Lopez-Ceron, M.; Cordova, H.; Sanchez-Montes, C.; Rodriguez de Miguel, C.; Sanchez, F.J. Exploring the clinical potential of an automatic colonic polyp detection method based on the creation of energy maps. Endoscopy 2016, 48, 837–842. [Google Scholar] [CrossRef] [PubMed]
  111. Bernal, J.; Sanchez, F.J.; Fernandez-Esparrach, G.; Gil, D.; Rodriguez, C.; Vilarino, F. WM-DOVA maps for accurate polyp highlighting in colonoscopy: Validation vs. saliency maps from physicians. Comput. Med. Imaging Graph. 2015, 43, 99–111. [Google Scholar] [CrossRef] [PubMed]
  112. Sehgal, V.; Rosenfeld, A.; Graham, D.G.; Banks, M.R.; Haidry, R.J.; Lovat, L.B. PTU-058 Machine Learning Creates A Simple Endoscopic Classification System for Detecting Dysplasia in Barrett’s Oesophagus with I-scan Imaging and Opens The Way To Standardised Training And Assessment Of Competence. Gut 2014, 63, A63. [Google Scholar] [CrossRef]
  113. Ughi, G.J.; Gora, M.J.; Swager, A.-F.; Soomro, A.; Grant, C.; Tiernan, A.; Rosenberg, M.; Sauk, J.S.; Nishioka, N.S.; Tearney, G.J. Automated segmentation and characterization of esophageal wall in vivo by tethered capsule optical coherence tomography endomicroscopy. Biomed. Opt. Express 2016, 7, 409–419. [Google Scholar] [CrossRef] [Green Version]
  114. Wang, Z.; Lee, H.-C.; Ahsen, O.O.; Liang, K.; Figueiredo, M.; Huang, Q.; Fujimoto, G.J.; Mashimo, H. Computer-Aided Analysis of Gland-Like Subsurface Hyposcattering Structures in Barrett’s Esophagus Using Optical Coherence Tomography. Appl. Sci. 2018, 8, 2420. [Google Scholar] [CrossRef]
  115. Kamboj, A.K.; Hoversten, P.; Kahn, A.K.; Trindade, A.J.; Iyer, P.G.; Wang, K.K.; Leggett, C.L. Interpretation of volumetric laser endomicroscopy in Barrett’s esophagus using image enhancement software. Dis. Esophagus 2019. [Google Scholar] [CrossRef]
  116. Lu, W.; Tong, Y.; Yu, Y.; Xing, Y.; Chen, C.; Shen, Y. Deep Learning-Based Automated Classification of Multi-Categorical Abnormalities from Optical Coherence Tomography Images. Transl. Vis. Sci. Technol. 2018, 7, 41. [Google Scholar] [CrossRef]
  117. Cai, S.; Zhong, Y.; Zhou, P.-H. Tu2005 development and validation of a deep learning algorithm for detection and recognition of precancerous lesion in esophagus. Gastrointest. Endosc. 2019, 89, AB654. [Google Scholar] [CrossRef]
  118. Shibata, J.; Ozawa, T.; Ishihara, S.; Onishi, T.; Matsuo, K.; Miura, M.; Aoyama, K.; Tada, T. Mo2050—Novel Computer-Assisted Detection System of Colorectal Carcinoid Tumors Using Convolutional Neural Networks. Gastroenterology 2019, 156, S-937. [Google Scholar] [CrossRef]
  119. Tokai, Y.; Yoshio, T.; Fujisaki, J.; Aoyama, K.; Tada, T. Sa1209 application of artificial intelligence using convolutional neural networks in diagnosing the invasion depth of esophageal squamous cell carcinoma. Gastrointest. Endosc. 2019, 89, AB169. [Google Scholar] [CrossRef]
  120. Nakagawa, K.; Ishihara, R.; Aoyama, K.; Kono, M.; Fukuda, H.; Shimamoto, Y.; Ohmori, M.; Iwagami, H.; Matsuno, K.; Inoue, S.; et al. Sa1257 specialist-level diagnostic accuracy for cancer invasion depth of superficial esophageal squamous cell carcinoma with conventional neural network. Gastrointest. Endosc. 2019, 89, AB191. [Google Scholar] [CrossRef]
  121. Xi, J.; Huo, L.; Wu, Y.; Cobb, M.J.; Hwang, J.H.; Li, X. High-resolution OCT balloon imaging catheter with astigmatism correction. Opt. Lett. 2009, 34, 1943–1945. [Google Scholar] [CrossRef] [PubMed]
  122. Su, J.; Zhang, J.; Yu, L.; Colt, H.G.; Brenner, M.; Chen, Z. Real-time swept source optical coherence tomography imaging of the human airway using a microelectromechanical system endoscope and digital signal processor. J. Biomed. Opt. 2008, 13, 030506. [Google Scholar] [CrossRef] [PubMed]
  123. Huo, L.; Xi, J.; Wu, Y.; Li, X. Forward-viewing resonant fiber-optic scanning endoscope of appropriate scanning speed for 3D OCT imaging. Opt. Express 2010, 18, 14375–14384. [Google Scholar] [CrossRef] [PubMed]
  124. Liang, K.; Traverso, G.; Lee, H.-C.; Ahsen, O.O.; Wang, Z.; Potsaid, B.; Giacomelli, M.; Jayaraman, V.; Barman, R.; Cable, A.; et al. Ultrahigh speed en face OCT capsule for endoscopic imaging. Biomed. Opt. Express 2015, 6, 1146–1163. [Google Scholar] [CrossRef] [Green Version]
  125. Iddan, G.; Meron, G.; Glukhovsky, A.; Swain, P. Wireless capsule endoscopy. Nature 2000, 405, 417. [Google Scholar] [CrossRef]
  126. Ramirez, F.C.; Akins, R.; Shaukat, M. Screening of Barrett’s esophagus with string-capsule endoscopy: A prospective blinded study of 100 consecutive patients using histology as the criterion standard. Gastrointest. Endosc. 2008, 68, 25–31. [Google Scholar] [CrossRef]
  127. Liao, Z.; Gao, R.; Xu, C.; Xu, D.-F.; Li, Z.-S. Sleeve string capsule endoscopy for real-time viewing of the esophagus: A pilot study (with video). Gastrointest. Endosc. 2009, 70, 201–209. [Google Scholar] [CrossRef] [PubMed]
  128. Gora, M.J.; Sauk, J.S.; Carruth, R.W.; Lu, W.; Carlton, D.T.; Soomro, A.; Rosenberg, M.; Nishioka, N.S.; Tearney, G.J. Imaging the Upper Gastrointestinal Tract in Unsedated Patients Using Tethered Capsule Endomicroscopy. Gastroenterology 2013, 145, 723–725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Endo, T.; Awakawa, T.; Takahashi, H.; Arimura, Y.; Itoh, F.; Yamashita, K.; Sasaki, S.; Yamamoto, H.; Tang, X.; Imai, K. Classification of Barrett’s epithelium by magnifying endoscopy. Gastrointest. Endosc. 2002, 55, 641–647. [Google Scholar] [CrossRef] [PubMed]
  130. Kudo, S.; Tamura, S.; Nakajima, T.; Yamano, H.; Kusaka, H.; Watanabe, H. Diagnosis of colorectal tumorous lesions by magnifying endoscopy. Gastrointest. Endosc. 1996, 44, 8–14. [Google Scholar] [CrossRef]
  131. Liang, C.P.; Dong, J.; Ford, T.; Reddy, R.; Hosseiny, H.; Farrokhi, H.; Beatty, M.; Singh, K.; Osman, H.; Vuong, B.; et al. Optical coherence tomography-guided laser marking with tethered capsule endomicroscopy in unsedated patients. Biomed. Opt. Express 2019, 10, 1207–1222. [Google Scholar] [CrossRef] [PubMed]
  132. Liang, K.; Wang, Z.; Ahsen, O.O.; Lee, H.-C.; Potsaid, B.M.; Jayaraman, V.; Cable, A.; Mashimo, H.; Li, X.; Fujimoto, J.G. Cycloid scanning for wide field optical coherence tomography endomicroscopy and angiography in vivo. Optica 2018, 5, 36–43. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) Time-domain optical coherence tomography (TD-OCT) uses low coherence interferometry. The light source is split and the backscattered light is captured by the detector. The depth information is provided by the scanned reference path. Adapted from Tsai et al. 2014. (b). Swept-source optical coherence tomography (SS-OCT) allows 10–200-fold faster imaging compared to TD-OCT. (A) Interferometer with swept laser source and beam splitter and path difference ΔL. (B) Time delay of light from sample (dotted) and reference light (dashed). (C) Interference signal proportional to delay. (D) Fourier transform of the interference signal measures ΔL. Figure and caption adapted from Tsai et al. 2014.
Figure 1. (a) Time-domain optical coherence tomography (TD-OCT) uses low coherence interferometry. The light source is split and the backscattered light is captured by the detector. The depth information is provided by the scanned reference path. Adapted from Tsai et al. 2014. (b). Swept-source optical coherence tomography (SS-OCT) allows 10–200-fold faster imaging compared to TD-OCT. (A) Interferometer with swept laser source and beam splitter and path difference ΔL. (B) Time delay of light from sample (dotted) and reference light (dashed). (C) Interference signal proportional to delay. (D) Fourier transform of the interference signal measures ΔL. Figure and caption adapted from Tsai et al. 2014.
Applsci 09 02991 g001
Figure 2. Combined en face and volumetric OCT dataset visualization. (a) Volumetric OCT dataset with cross-sections (blue), longitudinal (red) and en face (green) views. (b) Intrinsic co-registration leads to depth-resolved en face OCT images at various depths, revealing superficial mucosal architecture near the surface and mucosal patterns in greater detail and contrast, as well as microvasculature, in deeper images. Scale bars 1 mm. Figure and caption adapted from Ahsen et al. 2019.
Figure 2. Combined en face and volumetric OCT dataset visualization. (a) Volumetric OCT dataset with cross-sections (blue), longitudinal (red) and en face (green) views. (b) Intrinsic co-registration leads to depth-resolved en face OCT images at various depths, revealing superficial mucosal architecture near the surface and mucosal patterns in greater detail and contrast, as well as microvasculature, in deeper images. Scale bars 1 mm. Figure and caption adapted from Ahsen et al. 2019.
Applsci 09 02991 g002
Figure 3. Volumetric laser endomicroscopy (VLE) of the esophagus showing a luminal en face view of an area of overlap (yellow arrow) between the 3 features of dysplasia (orange is lack of layering, blue is glandular structures and pink is a hyper-reflective surface). (A) A view looking down from the proximal esophagus. (B) A view closer to the suspected area of dysplasia. The en face view is also shown (C). Figure and caption adapted from Trindade et al. 2019.
Figure 3. Volumetric laser endomicroscopy (VLE) of the esophagus showing a luminal en face view of an area of overlap (yellow arrow) between the 3 features of dysplasia (orange is lack of layering, blue is glandular structures and pink is a hyper-reflective surface). (A) A view looking down from the proximal esophagus. (B) A view closer to the suspected area of dysplasia. The en face view is also shown (C). Figure and caption adapted from Trindade et al. 2019.
Applsci 09 02991 g003

Share and Cite

MDPI and ACS Style

Samel, N.S.; Mashimo, H. Application of OCT in the Gastrointestinal Tract. Appl. Sci. 2019, 9, 2991. https://doi.org/10.3390/app9152991

AMA Style

Samel NS, Mashimo H. Application of OCT in the Gastrointestinal Tract. Applied Sciences. 2019; 9(15):2991. https://doi.org/10.3390/app9152991

Chicago/Turabian Style

Samel, Nicholas S., and Hiroshi Mashimo. 2019. "Application of OCT in the Gastrointestinal Tract" Applied Sciences 9, no. 15: 2991. https://doi.org/10.3390/app9152991

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