Next Article in Journal
Prenatal Diagnosis of Neu–Laxova Syndrome
Previous Article in Journal
Combinatorial Analysis of Circulating Biomarkers and Maternal Characteristics for Preeclampsia Prediction in the First and Third Trimesters in Asia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Editorial

Tissue Optical Clearing: State of the Art and Prospects

by
Elina A. Genina
1,2
1
Optics and Biophotonics Department, Saratov State University, 83 Astrakhanskaya Str., 410012 Saratov, Russia
2
Laboratory of Laser Molecular Imaging and Machine Learning, Tomsk State University, 36 Lenin’s Av., 634050 Tomsk, Russia
Diagnostics 2022, 12(7), 1534; https://doi.org/10.3390/diagnostics12071534
Submission received: 27 May 2022 / Accepted: 22 June 2022 / Published: 23 June 2022
(This article belongs to the Section Optical Diagnostics)
The term “tissue optical clearing” (TOC) came into use at the end of the 20th century and is associated with the development of methods for controlling tissue scattering properties using the refractive index matching effect. However, the first mention of increasing the transparency of biological tissues using immersion agents can be attributed to the work of Spalteholz in 1914 [1], in which an organic solvent-based technique was applied to tissue samples in vitro. The next step was made by Barer et al., in 1955 [2], who proposed the optical clearing of cell suspensions by means of protein solution with the same refractive index as the cell cytoplasm. The use of the TOC approach for an increase in eye sclera transparency was first applied by a scientific group with Tuchin in 1987 [3]. The authors demonstrated, theoretically and experimentally, the possibility of effective control of the scattering properties of sclera due to verografin solution. The works published from the mid-1990s by a number of leading scientific groups opened up a new direction in the study of TOC [4,5,6]. The rapid evolution of optical diagnostic and research methods induced comprehensive study of the mechanisms of TOC, a search for new approaches and optical clearing agents (OCAs). The development of understanding of the features and mechanisms of the phenomenon of TOC in vitro, ex vivo and in vivo demonstrated the method’s capabilities to increase the probing depth or the image contrast of optical devices, as reflected in reviews [7,8,9,10,11]. It was also found that the efficiency of TOC and the diffusion rate of OCAs differ in normal and pathological tissues that can be used as an additional diagnostic marker, for example, for diabetes [12] and cancer [13].
The development of fluorescence microscopy provided the possibility of obtaining a 3D reconstruction of the tissue anatomy and stimulated the development of innovative protocols of TOC and a new generation of OCAs for fixed samples. These protocols made it possible to almost completely overcome light scattering in tissues, organs or small animals and make them optically transparent materials. The FocusClear™ solution was one of the first specially designed OCAs, which was used by Chiang et al. in 2001 for the clearing of fixed insect organs with immunohistochemical labeling [14]. An alternative approach based on hyperhydration of tissue samples using the Scale technique was developed by Hama et al. in 2011 [15]. In 2013, hydrogel-tissue-hybridization called CLARITY was suggested by Chung et al. [16]. Detailed reviews are presented in Refs. [17,18,19] and book chapters [20].
Most soft tissues are characterized by a heterogeneous composition, including relatively large structures in the form of cells, fibers, lipid droplets and smaller components, such as cell organelles and phospholipid membranes. All of them are distributed in a water-based background matrix. Thus, tissues have a wide range of refractive indices, from ~1.34 for interstitial fluid and cytoplasm to >1.44 for proteins and lipids. The inorganic matrix of bones contains calcium hydroxyapatite with a refractive index >1.6. Refractive index mismatch of these components and the disordered packing of fibers in tissues are sources of strong light scattering. The main purpose of TOC is to match the refractive indices of the various tissue components using OCAs. However, different approaches reach this goal in different ways.
As a first approximation, the TOC process can be divided into three stages. The exchange flow of free water from the tissue to the OCA and the OCA to the tissue is characterized by different rates. Thus, at the first stage, hyperosmotic OCA causes a predominant loss of water in the tissues (dehydration). At the second stage, the diffusion of OCA in the tissue predominates, and at the third stage, the OCAs interact with the tissue components [21].
Molecular dynamics modeling has shown that the experimentally observed degree of TOC and parameters, such as osmolarity and refractive index, of OCA do not correlate. In Refs. [22,23,24], the hydrogen bond formation between different OCAs (alcohols and sugars) and collagen molecules was simulated. It has been shown that the higher conformational mobility of the low-molecular OCAs provides better adaptation of the agent to the collagen molecular pocket. OCA molecules partially displace bound water and disrupt the network of hydrogen bonds between collagen fibrils, which leads to swelling of fibrillar proteins at the third stage of TOC [24].
TOC mechanisms and approaches are discussed in detail in Ref. [21]. Briefly, TOC can be divided into several approaches depending on specific scientific problems: (i) simple immersion with the use of hyperosmotic or lipophilic OCAs, as one-component agents, as well as multi-component solutions; (ii) immersion with previous dehydration and delipidation with the use of organic solvents; (iii) hyperhydration with the use of delipidation and partial denaturation of tissue proteins; and (iv) hydrogel embedding with the use of fixation and crosslinking. The first two methods are associated with an increase in the average refractive index of tissue by replacing tissue water with a fluid with a higher refractive index. In contrast, the last two approaches involve removing lipids and reducing the refractive index of tissue samples to ~1.38. In hyperhydration, urea is used as a component for partial denaturation and hydration of hydrophobic regions of proteins. The fourth method allows fixing proteins and nucleic acids in their physiological locations by covalently linking the molecules to an acrylic-based hydrogel. Thus, the tissue samples can then be successfully cleared with OCAs, achieving high transparency. Some TOC protocols also allow for tissue decalcification and decolorization [21].
The first approach was successfully used in the development of in-vivo TOC. In these cases, TOC protocols should also provide for increasing the permeability of tissues and ensuring the safety and reversibility of the procedure.
In the last decade, a great variety of TOC methods based on different approaches was developed to achieve complete transparency of fixed tissue for fluorescence microscopy. Specially optimized multi-component solutions with innovative protocols, such as ACT-PRESTO, CLARITY, CRISTAL, CUBIC, DISCO, eFLASH, FlyClear, FRUIT, MOVIE, PEGASOS, Scale, SHANEL, SWITCH and many others, were suggested for application in visualization techniques [17,18,19,20,25]. They provided the effective clearing of tissues and whole organs, including bones. For example, vDISCO allowed 3D visualization of whole-body neuronal connectivity, meningeal lymphatic vessels and immune cells through the intact skull and vertebra in naive animals and trauma models [26]. An approach combining CLARITY with double-photon microscopy was proposed for the visualization of gray matter in the intact brain [27]. The CUBIC protocol was shown to be effective for an improvement in confocal imaging of mouse heart sections [28] and 3D vasculature images of a whole mouse heart [29]. A combination of CUBIC and CLARITY was used to evaluate the spatial distribution and phenotype of fibroblasts in mice left ventricles in a postmyocardial infarction [30]. A high-resolution whole mouse brain atlas was achieved by using CUBIC-X [31]. In order to better preserve tissue architecture and fluorescence, an alternative hydrogel-based method, named SHIELD, was used to evaluate the synaptic architecture of virally labeled murine neurons at single-cell resolution [32]. A hydrogel-based TOC method, MYOCLEAR, was developed for the labeling and detection of all neuromuscular junctions and diaphragm muscles in mice [33]. SHANEL protocol [34] was applied to large lipid-rich samples, such as human and porcine tissues, for clearing and decolorization.
TOC in vivo and ex vivo is realized using biocompatible OCAs separately or in combination with chemical and physical actions for enhancement of tissue permeability. In this case, biologically safe solvents in small concentrations can be included into the content of OCAs [35,36,37]. The use of DMSO, sonophoresis, laser irradiation, microdermabrasion, microperforation and other enhancers of OCA penetration in tissues are described in Refs. [38,39,40,41,42]. It is important for the development of these approaches to provide tissue viability. After the removal of OCAs, the tissue structure, functions and optical properties should be totally restored.
TOC protocols developed for skull clearing in vivo (SOCS, USOCA, SOCW, etc.) allowed for the visualization of blood vessels and brain tissues in mice without removing the cranial bone [36,43,44].
Despite significant progress in TOC, several important challenges remain. The variety of existing TOC methods and the constant work on the creation of new approaches indicate the impossibility of creating a universal protocol. The choice of the appropriate TOC protocol depends on tissue size and type, type of fluorescence, importance of tissue shrinkage and clearing time [17].
Until now, for many TOC approaches, the choice of OCA components and parameters for the additional enhancement of tissue permeability has been random and is not justified by the results of preliminary systematic studies. However, it is known that changes in the optical, weight and geometric parameters of tissue under the action of immersion fluids have significant concentration dependence [45]. Further, it was found that the efficiency of the subsequent clearing of fixed samples depends on the pH of fixation [46]. Taking into account such data would help to increase the efficiency of the development of new optical clearing methods. Molecular dynamics modeling can be useful to predict optimal OCAs.
Another important problem is the absorption of light by endogenous pigments, such as hemoglobin, myoglobin and melanin, which limits both the excitation light entering the tissue and the fluorescent radiation returning to the detector [17,21]. This requires the inclusion of additional bleaching components to TOC protocols [32,47].
Many organic solvents used cause a substantial dehydration and, thus, tissue shrinkage, as well as the quenching of fluorescent protein emissions [18]. To overcome these problems, it is necessary to complicate and prolong TOC protocols. Thus, sample preparation can take more than a month in some cases.
An example of the development of OCA based on chemical screening allowed expanding the palette of CUBIC approaches and the use of electrophoresis for delipidation and decolorization speeded up the procedure from a month to 2 weeks [48]. The MACS protocol based on a new hyperhydration reagent—M-xylylenediamine (MXDA)—possessed superior hyperhydration ability and showed much faster clearing than the Scale and CUBIC series methods [49].
Further refinements of the TOC methods contribute to whole-cell profiling of the human body. Therefore, the SHANEL method reported by Zhao at al. was applied to clear the entire human brain and kidney [34].
In in-vivo studies, a significant reduction in treatment time (no more than tens of minutes) and the use of only biologically compatible agents should be provided. It should also take into account the physiological response of living tissue to the action of the agent and its redistribution with a possible rapid exit from the area of interest. Therefore, the use of laser ablation and microporation in the stratum corneum ex vivo showed an increase in the penetration of OCA into the dermis; however, these treatments caused skin erythema and edema when applied in vivo, which increased both the absorption and scattering of light [40,50]. On the contrary, OCA in combination with sonophoresis caused rapid and safe optical clearing of the skin [41].
In parallel with the development of new TOC protocols, new microscopy technologies using clearing-specific objectives will follow. In addition to improved data collection approaches, the analysis and comparison of massive datasets is a relevant issue. Converting all human cells into digitized data is one of the ultimate goals of TOC technology [17,18,19,20,25]. Deep machine learning can be useful for getting statistically robust conclusions.
In the near future, the development of new methods will continue along with preclinical validation and the transfer of TOC approaches from the laboratory to clinical practice.

Funding

This research was funded by the RFBR grant No. 20-52-56005.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Spalteholz, W. Uber Das Durchsichtigmachen Von Menschlichen und Tierischen Praparaten und Seine Theoretischen Bedingungen; S. Hirzel: Leipzig, Germany, 1914. [Google Scholar]
  2. Barer, R. Spectrophotometry of clarified cell suspensions. Science 1955, 121, 709–715. [Google Scholar] [CrossRef] [PubMed]
  3. Bakutkin, V.V.; Maksimova, I.L.; Saprykin, P.I.; Tuchin, V.V.; Shubochkin, L.P. Light scattering by the human eye sclera. J. Appl. Spectrosc. 1987, 46, 104–107. [Google Scholar] [CrossRef]
  4. Chance, B.; Liu, H.; Kitai, T.; Zhang, Y. Effects of solutes on optical properties of biological materials: Models, cells, and tissues. Anal. Biochem. 1995, 227, 351–362. [Google Scholar] [CrossRef] [PubMed]
  5. Tuchin, V.V.; Maksimova, I.L.; Zimnyakov, D.A.; Kon, I.L.; Mavlutov, A.H.; Mishin, A.A. Light propagation in tissues with controlled optical properties. J. Biomed. Opt. 1997, 2, 401–417. [Google Scholar] [CrossRef]
  6. Vargas, G.; Chan, E.K.; Barton, J.K.; Rylander, H.G.; Welch, A.J. Use of an agent to reduce scattering in skin. Lasers Surg. Med. 1999, 24, 133–141. [Google Scholar] [CrossRef]
  7. Tuchin, V.V. A clear vision for laser diagnostics (Review). IEEE J. Sel. Top. Quantum Electron. 2007, 13, 1621–1628. [Google Scholar] [CrossRef]
  8. Genina, E.A.; Bashkatov, A.N.; Tuchin, V.V. Tissue optical immersion clearing. Expert Rev. Med. Devices 2010, 7, 825–842. [Google Scholar] [CrossRef]
  9. Zhu, D.; Larin, K.V.; Luo, Q.; Tuchin, V.V. Recent progress in tissue optical clearing. Laser Photonics Rev. 2013, 7, 732–757. [Google Scholar] [CrossRef] [Green Version]
  10. Sdobnov, A.Y.U.; Darvin, M.E.; Genina, E.A.; Bashkatov, A.N.; Lademann, J.; Tuchin, V.V. Recent progress in tissue optical clearing for spectroscopic application. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2018, 197, 216–229. [Google Scholar] [CrossRef]
  11. Costantini, I.; Cicchi, R.; Silvestri, L.; Vanzi, F.; Pavone, F.S. In-Vivo and Ex-Vivo optical clearing methods for biological tissues: Review. Biomed. Opt. Express 2019, 10, 5251–5267. [Google Scholar] [CrossRef]
  12. Tuchina, D.K.; Shi, R.; Bashkatov, A.N.; Genina, E.A.; Zhu, D.; Luo, Q.; Tuchin, V.V. Ex Vivo diffusion kinetics of glucose in native and in vivo glycated mouse skin. J. Biophotonics 2015, 8, 332–346. [Google Scholar] [CrossRef] [PubMed]
  13. Genina, E.A.; Oliveira, L.M.C.; Bashkatov, A.N.; Tuchin, V.V. Optical Clearing of Biological Tissues: Prospects of Application for Multimodal Malignancy Diagnostics. In Multimodal Optical Diagnostics of Cancer; Tuchin, V.V., Popp, J., Zakharov, V., Eds.; Springer Nature: Cham, Switzerland, 2020; pp. 107–132. [Google Scholar] [CrossRef]
  14. Chiang, A.S.; Liu, Y.C.; Chiu, S.L.; Hu, S.H.; Huang, C.Y.; Hsieh, C.H. Three-dimensional mapping of brain neuropils in the cockroach, Diploptera punctata. J. Comp. Neurol. 2001, 440, 1–11. [Google Scholar] [CrossRef] [PubMed]
  15. Hama, H.; Kurokawa, H.; Kawano, H.; Ando, R.; Shimogori, T.; Noda, H.; Fukami, K.; Sakaue-Sawano, A.; Miyawaki, A. Scale: A chemical approach for fluorescence imaging and reconstruction of transparent mouse brain. Nat. Neurosci. 2011, 14, 1481–1488. [Google Scholar] [CrossRef] [PubMed]
  16. Chung, K.; Wallace, J.; Kim, S.Y.; Kalyanasundaram, S.; Andalman, A.S.; Davidson, T.J.; Mirzabekov, J.J.; Zalocusky, K.A.; Mattis, J.; Denisin, A.K.; et al. Structural and molecular interrogation of intact biological systems. Nature 2013, 497, 332–337. [Google Scholar] [CrossRef] [PubMed]
  17. Richardson, D.S.; Lichtman, J.W. Clarifying tissue clearing. Cell 2015, 162, 246–257. [Google Scholar] [CrossRef] [Green Version]
  18. Matryba, P.; Kaczmarek, L.; Gołąb, J. Advances in Ex Situ tissue optical clearing. Laser Photonics Rev. 2019, 13, 1800292. [Google Scholar] [CrossRef] [Green Version]
  19. Gomez-Gaviro, M.V.; Sanderson, D.; Ripoll, J.; Desco, M. Biomedical applications of tissue clearing and three-dimensional imaging in health and disease. Iscience 2020, 23, 101432. [Google Scholar] [CrossRef]
  20. Tuchin, V.; Zhu, D.; Genina, E.A. (Eds.) Handbook of Tissue Optical Clearing: New Prospects in Optical Imaging; CRC Press: Boca Raton, FL, USA, 2022; 658p. [Google Scholar] [CrossRef]
  21. Yu, T.; Zhu, D.; Oliveira, L.; Genina, E.A.; Bashkatov, A.N.; Tuchin, V.V. Tissue optical clearing mechanisms. In Handbook of Tissue Optical Clearing: New Prospects in Optical Imaging; Tuchin, V., Zhu, D., Genina, E.A., Eds.; CRC Press: Boca Raton, FL, USA, 2022; pp. 3–30. [Google Scholar] [CrossRef]
  22. Yeh, A.T.; Hirshburg, J. Molecular interactions of exogenous chemical agents with collagen—Implications for tissue optical clearing. J. Biomed. Opt. 2006, 11, 014003. [Google Scholar] [CrossRef] [Green Version]
  23. Feng, W.; Shi, R.; Ma, N.; Tuchina, D.K.; Tuchin, V.V.; Zhu, D. Skin optical clearing potential of disaccharides. J. Biomed. Opt. 2016, 21, 081207. [Google Scholar] [CrossRef] [Green Version]
  24. Berezin, K.V.; Dvoretskiy, K.N.; Chernavina, M.L.; Likhter, A.M.; Smirnov, V.V.; Shagautdinova, I.T.; Antonova, E.M.; Stepanovich, E.Y.; Dzhalmuhambetova, E.A.; Tuchin, V.V. Molecular modeling of immersion optical clearing of biological tissues. J. Mol. Model. 2018, 24, 45. [Google Scholar] [CrossRef]
  25. Susaki, E.A.; Ueda, H.R. Challenges and opportunities in hydrophilic tissue clearing methods. In Handbook of Tissue Optical Clearing: New Prospects in Optical Imaging; Tuchin, V., Zhu, D., Genina, E.A., Eds.; CRC Press: Boca Raton, FL, USA, 2022; pp. 257–276. [Google Scholar] [CrossRef]
  26. Cai, R.; Pan, C.; Ghasemigharagoz, A.; Todorov, M.; Foerstera, B.; Zhao, S.; Bhatia, H.S.; Mrowka, L.; Theodorou, D.; Rempfler, M.; et al. Panoptic vDISCO imaging reveals neuronal connectivity, remote trauma effects and meningeal vessels in intact transparent mice. BioRxiv 2018, 374785. [Google Scholar] [CrossRef]
  27. Chang, E.H.; Argyelan, M.; Aggarwal, M.; Chandon, T.-S.S.; Karlsgodt, K.H.; Mori, S.; Malhotra, A.K. The role of myelination in measures of white matter integrity: Combination of diffusion tensor imaging and two-photon microscopy of CLARITY intact brains. NeuroImage 2017, 147, 253–261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Nehrhoff, I.; Bocancea, D.; Vaquero, J.; Vaquero, J.J.; Ripoll, J.; Desco, M.; Gómez-Gaviro, M.V. 3D imaging in CUBIC-cleared mouse heart tissue: Going deeper. Biomed. Opt. Express 2016, 29, 3716–3720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Nehrhoff, I.; Ripoll, J.; Samaniego, R.; Desco, M.; Gomez-Gaviro, M.V. Looking inside the heart: A see-through view of the vascular tree. Biomed. Opt. Express 2017, 8, 3110–3118. [Google Scholar] [CrossRef] [PubMed]
  30. Wang, Z.; Zhang, J.; Fan, G.; Zhao, H.; Wang, X.; Zhang, J.; Zhang, P.; Wang, W. Imaging transparent intact cardiac tissue with single-cell resolution. Biomed. Opt. Express 2018, 9, 423–436. [Google Scholar] [CrossRef]
  31. Murakami, T.C.; Mano, T.; Saikawa, S.; Horiguchi, S.A.; Shigeta, D.; Baba, K.; Sekiya, H.; Shimizu, Y.; Tanaka, K.F.; Kiyonari, H.; et al. A three-dimensional single-cell-resolution whole-brain atlas using CUBIC-X expansion microscopy and tissue clearing. Nat. Neurosci. 2018, 21, 625–637. [Google Scholar] [CrossRef]
  32. Pende, M.; Vadiwala, K.; Schmidbaur, H.; Stockinger, A.W.; Murawala, P.; Saghafi, S.; Dekens, M.P.S.; Becker, K.; Revilla-I-Domingo, R.; Papadopoulos, S.-C.; et al. A versatile depigmentation, clearing, and labeling method for exploring nervous system diversity. Sci. Adv. 2020, 6, eaba0365. [Google Scholar] [CrossRef]
  33. Williams, M.P.I.; Rigon, M.; Straka, T.; Hörner, S.J.; Thiel, M.; Gretz, N.; Hafner, M.; Reischl, M.; Rudolf, R. A novel optical tissue clearing protocol for mouse skeletal muscle to visualize endplates in their tissue context. Front. Cell Neurosci. 2019, 13, 49. [Google Scholar] [CrossRef] [Green Version]
  34. Zhao, S.; Todorov, M.I.; Cai, R.; Ai-Maskari, R.; Steinke, H.; Kemter, E.; Mai, H.; Rong, Z.; Warmer, M.; Stanic, K.; et al. Cellular and molecular probing of intact human organs. Cell 2020, 180, 796–812.e19. [Google Scholar] [CrossRef]
  35. Zhao, Q.; Dai, C.; Fan, S.; Lv, J.; Nie, L. Synergistic efficacy of salicylic acid with a penetration enhancer on human skin monitored by OCT and diffuse reflectance spectroscopy. Sci. Rep. 2016, 6, 34954. [Google Scholar] [CrossRef]
  36. Feng, W.; Zhang, C.; Yu, T.; Semyachkina-Glushkovskaya, O.; Zhu, D. In Vivo monitoring blood–brain barrier permeability using spectral imaging through optical clearing skull window. J. Biophotonics 2019, 12, e201800330. [Google Scholar] [CrossRef] [PubMed]
  37. Zaytsev, S.M.; Amouroux, M.; Khairallah, G.; Bashkatov, A.N.; Tuchin, V.V.; Blondel, W.; Genina, E.A. Impact of optical clearing on ex vivo human skin optical properties characterized by spatially resolved autofluorescence and diffuse reflectance spectroscopy. J. Biophotonics 2022, 15, e202100202. [Google Scholar] [CrossRef] [PubMed]
  38. Liu, C.; Zhi, Z.; Tuchin, V.V.; Luo, Q.; Zhu, D. Enhancement of skin optical clearing efficacy using photo-irradiation. Lasers Surg. Med. 2010, 42, 132–140. [Google Scholar] [CrossRef] [PubMed]
  39. Damestani, Y.; Melakeberhan, B.; Rao, M.P.; Aguilar, G. Optical clearing agent perfusion enhancement via combination of microneedle poration, heating and pneumatic pressure. Lasers Surg. Med. 2014, 46, 488–498. [Google Scholar] [CrossRef] [Green Version]
  40. Genina, E.A.; Bashkatov, A.N.; Terentyuk, G.S.; Tuchin, V.V. Integrated effects of fractional laser microablation and sonophoresis on skin immersion optical clearing In Vivo. J. Biophotonics 2020, 13, e202000101. [Google Scholar] [CrossRef]
  41. Genina, E.A.; Surkov, Y.I.; Serebryakova, I.A.; Bashkatov, A.N.; Tuchin, V.V.; Zharov, V.P. Rapid ultrasound optical clearing of human light and dark skin. IEEE Trans. Med. Imaging 2020, 39, 3198–3206. [Google Scholar] [CrossRef]
  42. Yanina, I.Y.; Tanikawa, Y.; Genina, E.A.; Dyachenko, P.A.; Tuchina, D.K.; Bashkatov, A.N.; Dolotov, L.E.; Tarakanchikova, Y.V.; Terentuk, G.S.; Navolokin, N.A.; et al. Immersion optical clearing of adipose tissue in rats: Ex Vivo and In Vivo studies. J. Biophotonics 2022, e202100393. [Google Scholar] [CrossRef]
  43. Wang, J.; Zhang, Y.; Xu, T.; Luo, Q.; Zhu, D. An innovative transparent cranial window based on skull optical clearing. Laser Phys. Lett. 2012, 9, 469–473. [Google Scholar] [CrossRef]
  44. Zhang, C.; Feng, W.; Zhao, Y.J.; Yu, T.; Li, P.; Xu, T.; Luo, Q.; Zhu, D. A large, switchable optical clearing skull window for cerebrovascular imaging. Theranostics 2018, 8, 2696–2708. [Google Scholar] [CrossRef]
  45. Genin, V.D.; Genina, E.A.; Tuchin, V.V.; Bashkatov, A.N. Glycerol effects on optical, weight and geometrical properties of skin tissue. J. Innov. Opt. Health Sci. 2021, 14, 2142006. [Google Scholar] [CrossRef]
  46. Susaki, E.A.; Tainaka, K.; Perrin, D.; Yukinaga, H.; Kuno, A.; Ueda, H.R. Advanced CUBIC protocols for whole-brain and whole-body clearing and imaging. Nat. Protoc. 2015, 10, 1709–1727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Jing, D.; Zhang, S.; Luo, W.; Gao, X.; Men, Y.; Ma, C.; Liu, X.; Yi, Y.; Bugde, A.; Zhou, B.O.; et al. Tissue clearing of both hard and soft tissue organs with the PEGASOS method. Cell Res. 2018, 28, 803–818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Susaki, E.A.; Ueda, H.R. Whole-body and Whole-Organ Clearing and Imaging Techniques with Single-Cell Resolution: Toward Organism-Level Systems Biology in Mammals. Cell Chem. Biol. 2016, 23, 137–157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Zhu, J.; Yu, T.; Li, Y.; Xu, J.; Qi, Y.; Yao, Y.; Ma, Y.; Wan, P.; Chen, Z.; Li, X.; et al. MACS: Rapid aqueous clearing system for 3D mapping of intact organs. Adv. Sci. Lett. 2020, 7, 1903185. [Google Scholar] [CrossRef] [Green Version]
  50. Genina, E.A.; Ksenofontova, N.S.; Bashkatov, A.N.; Terentyuk, G.S.; Tuchin, V.V. Study of the epidermis ablation effect on the efficiency of optical clearing of skin In Vivo. Quant. Electron. 2017, 47, 561–566. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Genina, E.A. Tissue Optical Clearing: State of the Art and Prospects. Diagnostics 2022, 12, 1534. https://doi.org/10.3390/diagnostics12071534

AMA Style

Genina EA. Tissue Optical Clearing: State of the Art and Prospects. Diagnostics. 2022; 12(7):1534. https://doi.org/10.3390/diagnostics12071534

Chicago/Turabian Style

Genina, Elina A. 2022. "Tissue Optical Clearing: State of the Art and Prospects" Diagnostics 12, no. 7: 1534. https://doi.org/10.3390/diagnostics12071534

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