Open Access
16 November 2017 High-refractive index of acrylate embedding resin clarifies mouse brain tissue
Hongfu Zhou, Yumiao Xiong, Yu Wang, Xiaojun Wang, Pei Li, Yadong Gang, Xiuli Liu, Shaoqun Zeng
Author Affiliations +
Abstract
Biological tissue transparency combined with light-sheet fluorescence microscopy is a useful method for studying the neural structure of biological tissues. The development of light-sheet fluorescence microscopy also promotes progress in biological tissue clearing methods. The current clarifying methods mostly use liquid reagent to denature protein or remove lipids first, to eliminate or reduce the scattering index or refractive index of the biological tissue. However, denaturing protein and removing lipids require complex procedures or an extended time period. Therefore, here we have developed acrylate resin with a high refractive index, which causes clearing of biological tissue directly after polymerization. This method can improve endogenous fluorescence retention by adjusting the pH value of the resin monomer.

Biological tissue transparency combined with light-sheet fluorescent microscopy has become an important method for studying the neural structure of biological tissue, and methods for clarifying biological tissue have advanced greatly.13 Biological tissue transparency is mainly attained via denaturation of protein or removal of lipids of the cell membrane, to eliminate or reduce the refractive index or the scattering index of the biological tissue, so that the deeper organizational structure of nerves can be observed.4,5 The recently developed clarifying methods can be divided into three categories: aqueous-based clearing, hydrogel embedding combined with electrophoresis clearing, and solvent-based clearing. Aqueous-based clearing of biological tissues tends to cause tissue expansion and requires an extended period of time, and some denaturing reagents, such as urea or sodium dodecyl sulfate, could cause the tissue to become antigen-insensitive.613 Hydrogel embedding combined with electrophoresis clearing is a complex method that requires gel embedding and electrophoresis equipment, and the reproducibility of experimental results is poor.1416 Solvent-based reagents virtually quench endogenous fluorescent proteins.1720 Although the ultimate three-dimensional imaging of solvent cleared organs technique can clarify biological tissues expressing fluorescent proteins, the fluorescence intensity gradually diminishes after clearing.21 To resolve these problems, the fluorescent signal requires long-term preservation and antibleaching treatment after resin embedding; this method involves first using solvent-based clearing of the biological tissue, after which it is embedded in epoxy resin.22

We developed an acrylic resin with a high refractive index resin (termed HRAR) that clears biological tissue directly after embedding. The refractive index is increased when the liquid monomer polymerizes into solid resin, which matches the refractive index of the dehydrated biological tissue perfectly, to achieve transparency. This method can clear large samples directly without requiring special pretreatment, and can maintain the endogenous fluorescent signal well. In this study, we investigated the effect of differences in the pH value of monomers on green fluorescent protein (GFP) fluorescence,23 and a transparent GFP mouse brain block imaged via light-sheet fluorescence microscopy.

Monomer 2-phenoxyethyl acrylate and the crosslinking agent ethylene glycol dimethacrylate (mass ratio of 91) were polymerized in the presence of 0.2% initiator 2,2′-azobis(2,4-dimethyl) valeronitrile to produce HRAR resin (Fig. 1). The main effect of 2-phenoxyethyl acrylate is to increase the refractive index after polymerization, and the crosslinking agent triethylene glycol dimethacrylate causes the resin to form a crosslinked network structure that improves the hardness of the resin. The results (Fig. 2) showed that the light transmittance of an HM20 resin-embedded mouse brain was about 7% in the spectral range of 400 to 1100 nm, while the transmittance of an HRAR resin-embedded mouse brain was about 70% in this spectral range.

Fig. 1

The polymerization of HRAR. The structure (a) is monomer 2-phenoxyethyl acrylate, the structure (b) is crosslinking agent ethylene glycol dimethacrylate, and the protein structure represents the embedded biological tissue.

JBO_22_11_110503_f001.png

Fig. 2

The clearing effect of HRAR. (a) HM 20 resin-embedded mouse brain block; (b) HRAR resin-embedded mouse brain block; (c) optical transmittance of HM20 resin- and HRAR resin-embedded mouse brain blocks. A PerkinElmer (Waltham, Massachusetts) Lambda 35 UV/vis spectrometer was used at an excitation bandwidth of 0.5 nm, with a spectral range of 400 to 1100 nm was used to assess transmittance of an HM20 resin block and an HRAR resin block with 3-mm thickness.

JBO_22_11_110503_f002.png

Eight-week-old mice were perfused according to the recommended protocol for the perfusion of mice.24 Animal care and use were in accordance with the guidelines of the Administration Committee of Affairs Concerning Experimental Animals in Hubei Province of China. The protocol was approved by the Committee on the Ethics of Animal Experiments of the Huazhong University of Science and Technology (permit number: 00027340). All efforts were made to minimize the suffering of the animals. The tissue was sectioned into 100-μm coronal slices using a vibrating microtome (Leica, VT1000 S), and the slices were stored in 0.01 M PBS solution at 4°C in the dark.

The 100-μm brain slices were dehydrated according to the following procedure: 75% tetrahydrofuran (THF) for 3 min, 100% THF for 3 min, and 100% THF for 4 min. The brain slices were then placed in the resin mixture twice for 5 min each time. Two drops of resin mixture were then added to the slides, which were subsequently coverslipped. The brain slices on the slides were polymerized in an oven at 45°C for 8 h. This experiment was carried out to verify the endogenous fluorescence intensity before and after embedding in the HRAR resin. First, 100-μm brain slices were placed on a slide and coverslipped. Confocal fluorescence microscopy (Zeiss 780) was then used to image the region of interest. The imaged slices were embedded according to the above procedure, and confocal fluorescence microscopy (Zeiss 780) was again used to image the same region using the same parameters. To quantitatively evaluate the endogenous fluorescence intensity of the tissue before and after embedding in the HRAR resin, we used the fluorescence intensity of the soma as an indicator. First, the imaging data were imported into Image J software. Within this software, the rectangular-selection tool was activated, a circular area on the soma was selected, and the histogram tool was then used to measure the average gray value; the average gray value of the soma before embedding was denoted as A, while the average gray value of the soma after embedding was denoted as B. The fluorescence retention rate after embedding was calculated as follows: B/A×100%. For each group of experiments, we selected 10 somata to evaluate the average gray values.

2-phenoxyethyl acrylate and ethylene glycol dimethacrylate were filtered through alkaline alumina (Al2O3), respectively, solution a was prepared by mixing 2-phenoxyethyl acrylate and ethylene glycol dimethacrylate in a mass ratio of 91, then 0.2% initiator 2,2′-azobis(2,4-dimethyl) valeronitrile was added, at last the mixture was stirred for 30 min; solution b was prepared by mixing 2-phenoxyethyl acrylate and ethylene glycol dimethacrylate in a mass ratio of 91, 10% tetrasodium ethylenediaminetetraacetate dehydrate (EDTA-Na4) was then added and the solution was shaken for 6 h. The mixture was filtered and then 0.2% initiator 2,2′-azobis(2,4-dimethyl) valeronitrile was added, at last the mixture was stirred for 30 min. Solution c was prepared by mixing 2-phenoxyethyl acrylate and ethylene glycol dimethacrylate in a mass ratio of 91, 10% tetrasodium ethylenediaminetetraacetate dehydrate (EDTA-Na4) was then added and the solution was shaken for 6 h. The mixture was filtered and then 0.2% initiator 2,2′-azobis(2,4-dimethyl) valeronitrile and 0.5‰ 2,2′-diaminodiethylamine were added, at last the mixture was stirred for 30 min. The three solutions that were measured with a pH meter (OHAUS S-5000 pH meter, Parsippany, New Jersey) were calculated the average pH value after multiple measurements. The results (Fig. 3) show that the fluorescence retention ratio of endogenous GFP signal was 46% when the pH value of the resin monomer was 4.81, 162% when the pH value of the resin monomer was 6.27, and 210% when the pH value of the resin monomer was 9.62. This indicates that the pH value of the resin monomer affected the fluorescence signal intensity markedly, and that an alkaline environment is beneficial to the maintenance of the endogenous GFP signal.

Fig. 3

The fluorescence comparison of GFP in brain slices after embedding in resin with different pH values. (a), (b), and (c) The imaging results of the brain slices in PBS buffer. (d) The imaging results after embedding the tissue in HRAR liquid agent at a pH of 4.81. (g) The fluorescence intensity distribution curve for the area indicated by colored lines in (a) and (d). (e) The imaging results after embedding tissue in HRAR liquid agent at a pH of 6.27. (h) The fluorescence intensity distribution curve for the area indicated by colored lines in (b) and (e). (f) The imaging results after embedding in HRAR liquid agent at a pH of 9.62. (i) The fluorescence intensity distribution curve for the area indicated by colored lines in (c) and (f). (j) The fluorescence intensity ratio of GFP before and after embedding of tissue in HRAR resin; the pH of column 1 is 4.81, the pH of column 2 is 6.27, and the pH of column 3 is 9.62, and the scale bar for all images (a)–(f) indicates 30  μm. As the tissue block shrink around 20% in each direction after embedding, (d)–(f) enlarged accordingly to make the soma be close to that of the original image. Therefore, the scale bar of (d)–(f) is longer than that in (a)–(c). For (a)–(f), three-dimensional images were obtained with tissue thickness of 100  μm, at a z-step of 1  μm, using a confocal fluorescence microscopy (Zeiss 780), equipped with a 488-nm laser, a 20× objective lens with numerical aperture of 1.0, zoom of 1.0, medium is water. (a)–(f) z-axis maximum projection images of these images stacks.

JBO_22_11_110503_f003.png

The postfixed mouse brain block was rinsed three times with 0.01 M PBS solution for 8 h each at 4°C in the dark, after which it was dehydrated in 50% THF for 2 h, 75% THF for 2 h, 95% THF for 2 h, 100% THF for 3 h, and 100% THF for 3 h at 4°C. The dehydrated brain block was then embedded in HRAR resin as follows: HRAR resin for 2 h, HRAR resin for 24 h, and HRAR resin for 48 h at 4°C in the dark. Next, the permeated mouse brain block was placed in a capsule that was filled with HRAR resin. The curing procedure was as follows: 32°C for 8 h, 38°C for 8 h, 45°C for 8 h, and 50°C for 4 h in vacuum and dark conditions. The results (Fig. 4) from light-sheet fluorescence microscopy imaging showed that the fluorescence signal and neuronal fiber structure were well preserved with the HRAR resin (solution c), and that neuronal fibers were not distorted by this embedding process. Thus, the polymerization rate was uniform, and this method could be applied for clearing large samples.

Fig. 4

(a) Light-sheet fluorescence microscopy imaging of an HRAR resin-embedded GFP brain block. (b) and (c) Magnified cortex nerve three-dimensional structure. Light-sheet fluorescence microscope (Ultramicroscope, LaVision BioTec, Bielefeld, Germany), equipped with a 488-nm laser, 2× objective lens with numerical aperture of 0.5, zoom of 1.6, in dibenzyl ether medium.

JBO_22_11_110503_f004.png

In this paper, we applied the principle of refractive index matching to clear biological tissue, and found that the refractive index increased by about 0.05 after the acrylic monomer polymerized. We then developed an acrylate resin with a high refractive index that could clear dehydrated biological tissue directly. The average refractive index of the resin monomer is 1.512, and the biological tissue was translucent after penetration by the resin monomer. The average refractive index of the polymerized resin is 1.558, and the refractive index of the dehydrated biological tissue is 1.55 to 1.56,1 the refractive index of the acrylate resin, therefore, matches that of the dehydrated biological tissue, which results in direct clearing of the biological tissue after polymerization. The refractive index of a traditional commercial acrylic monomer is about 1.40 to 1.45, and the refractive index of the solid commercial resin is about 1.50. Therefore, biological tissue embedded in such resin cannot be transparent, as the refractive index of the commercial resin is mismatched to that of the dehydrated biological tissue.

In contrast with traditional commercial acrylate resin, the HRAR resin greatly improved the ability of light to penetrate biological tissue. The light transmittance in the spectral range of 400 to 1100 nm does not change markedly, which ensured that standard light wavelengths could penetrate the embedded biological tissue, greatly reducing attenuation of the excitation and emission light.

We also investigated the fluorescence retention by adjusting the pH value of the resin monomer, and found that an alkaline environment is beneficial to maintaining endogenous GFP signals, while an acidic environment could quench endogenous GFP signals.23 Thus, the pH value of the resin monomer can affect endogenous GFP signal intensity markedly. In our study, solution c was the best for maintaining the GFP signal. Light-sheet microscopy imaging of brain block showed that the cell bodies and nerve fibers were maintained intact, indicating that this method is also suitable for large sample clearing and light-sheet fluorescent microscopy imaging.

Disclosures

The authors have no relevant financial interests in this article and no potential conflicts of interest to disclose.

Acknowledgments

We thank Jianyi Xu, Fang Fang Yin, Yadong Gang, Pei Li, Ke Bai, Lin Li, and Yuanlei Yue for their helpful comments. This work was supported by the NSFC Project Nos. 81327802 and 61421064.

References

1. 

D. S. Richardson and J. W. Lichtman, “Clarifying tissue clearing,” Cell, 162 (2), 246 –257 (2015). http://dx.doi.org/10.1016/j.cell.2015.06.067 CELLB5 0092-8674 Google Scholar

2. 

A. Feuchtinger, A. Walch and M. Dobosz, “Deep tissue imaging: a review from a preclinical cancer research perspective,” Histochem. Cell Biol., 146 (6), 781 –806 (2016). http://dx.doi.org/10.1007/s00418-016-1495-7 Google Scholar

3. 

B.-C. Chen et al., “Lattice light-sheet microscopy: imaging molecules to embryos at high spatiotemporal resolution,” Science, 346 (6208), 1257998 (2015). http://dx.doi.org/10.1126/science.1257998 SCIEAS 0036-8075 Google Scholar

4. 

H. Liu et al., “Dependence of tissue optical properties on solute-induced changes in refractive index and osmolarity,” J. Biomed. Opt., 1 (2), 200 –211 (1996). http://dx.doi.org/10.1117/12.231370 JBOPFO 1083-3668 Google Scholar

5. 

S. L. Jacques, “Corrigendum: optical properties of biological tissues: a review,” Phys. Med. Biol., 58 (14), 5007 –5008 (2013). http://dx.doi.org/10.1088/0031-9155/58/14/5007 PHMBA7 0031-9155 Google Scholar

6. 

Y.-C. Liu and A.-S. Chiang, “High-resolution confocal imaging and three-dimensional rendering,” Methods, 30 (1), 86 –93 (2003). http://dx.doi.org/10.1016/S1046-2023(03)00010-0 MTHDE9 1046-2023 Google Scholar

7. 

P. S. Tsai et al., “Correlations of neuronal and microvascular densities in murine cortex revealed by direct counting and colocalization of nuclei and vessels,” J. Neuroscience, 29 (46), 14553 –14570 (2009). http://dx.doi.org/10.1523/JNEUROSCI.3287-09.2009 Google Scholar

8. 

H. Hama et al., “Scale: a chemical approach for fluorescence imaging and reconstruction of transparent mouse brain,” Nat. Neurosci., 14 (11), 1481 –1488 (2011). http://dx.doi.org/10.1038/nn.2928 NANEFN 1097-6256 Google Scholar

9. 

T. Kuwajima et al., “ClearT: a detergent and solvent-free clearing method for neuronal and non-neuronal tissue,” Development, 140 (6), 1364 –1368 (2013). http://dx.doi.org/10.1242/dev.091844 Google Scholar

10. 

K. Tainaka et al., “Whole-body imaging with single-cell resolution by tissue decolorization,” Cell, 159 (4), 911 –924 (2014). http://dx.doi.org/10.1016/j.cell.2014.10.034 CELLB5 0092-8674 Google Scholar

11. 

B. Yang et al., “Single-cell phenotyping within transparent intact tissue through whole-body clearing,” Cell, 158 (4), 945 –958 (2014). http://dx.doi.org/10.1016/j.cell.2014.07.017 CELLB5 0092-8674 Google Scholar

12. 

B. Hou et al., “Scalable and DiI-compatible optical clearance of the mammalian brain,” Front. Neuroanat., 9 1 –19 (2015). http://dx.doi.org/10.3389/fnana.2015.00019 Google Scholar

13. 

Y. Aoyagi et al., “A rapid optical clearing protocol using 2,2′-thiodiethanol for microscopic observation of fixed mouse brain,” PLoS One, 10 (1), e0116280 (2015). http://dx.doi.org/10.1371/journal.pone.0116280 POLNCL 1932-6203 Google Scholar

14. 

K. Chung et al., “Structural and molecular interrogation of intact biological systems,” Nature, 497 (7449), 332 –337 (2013). http://dx.doi.org/10.1038/nature12107 Google Scholar

15. 

S.-Y. Kima et al., “Stochastic electrotransport selectively enhances the transport of highly electromobile molecules,” Proc. Natl. Acad. Sci. U.S.A., 112 (46), E6274 –E6283 (2015). http://dx.doi.org/10.1073/pnas.1510133112 Google Scholar

16. 

E. Lee et al., “ACT-PRESTO: rapid and consistent tissue clearing and labeling method for 3-dimensional (3D) imaging,” Sci. Rep., 6 18631 (2016). http://dx.doi.org/10.1038/srep18631 SRCEC3 2045-2322 Google Scholar

17. 

W. Spalteholz, Uber das Durchsichtigmachen von menschlichen und tierischen Praparaten, S. Hierzel, Leipzig (1914). Google Scholar

18. 

H.-U. Dodt et al., “Ultramicroscopy: three-dimensional visualization of neuronal networks in the whole mouse brain,” Nat. Methods, 4 (4), 331 –336 (2007). http://dx.doi.org/10.1038/nmeth1036 1548-7091 Google Scholar

19. 

A. Ertürk et al., “Three-dimensional imaging of solvent-cleared organs using 3DISCO,” Nat. Protoc., 7 (11), 1983 –1995 (2012). http://dx.doi.org/10.1038/nprot.2012.119 1754-2189 Google Scholar

20. 

N. Renier et al., “iDISCO: a simple, rapid method to immunolabel large tissue samples for volume imaging,” Cell, 159 (4), 896 –910 (2014). http://dx.doi.org/10.1016/j.cell.2014.10.010 CELLB5 0092-8674 Google Scholar

21. 

C. Pan et al., “Shrinkage-mediated imaging of entire organs and organisms using uDISCO,” Nat. Methods, 13 (10), 859 –867 (2016). http://dx.doi.org/10.1038/nmeth.3964 1548-7091 Google Scholar

22. 

K. Becker et al., “Reduction of photo bleaching and long term archiving of chemically cleared GFP-expressing mouse brains,” PLoS One, 9 (12), e114149 (2014). http://dx.doi.org/10.1371/journal.pone.0114149 POLNCL 1932-6203 Google Scholar

23. 

Z. Yang et al., “Development of a plastic embedding method for large volume and fluorescent-protein-expressing tissues,” PLoS One, 8 (4), e60877 (2013). http://dx.doi.org/10.1371/journal.pone.0060877 POLNCL 1932-6203 Google Scholar

24. 

Y. Gang et al., “Embedding and chemical reactivation of green fluorescent protein in the whole mouse brain for optical micro-imaging,” Front. Neurosci., 11 121 (2017). http://dx.doi.org/10.3389/fnins.2017.00121 1662-453X Google Scholar
CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Hongfu Zhou, Yumiao Xiong, Yu Wang, Xiaojun Wang, Pei Li, Yadong Gang, Xiuli Liu, and Shaoqun Zeng "High-refractive index of acrylate embedding resin clarifies mouse brain tissue," Journal of Biomedical Optics 22(11), 110503 (16 November 2017). https://doi.org/10.1117/1.JBO.22.11.110503
Received: 5 July 2017; Accepted: 6 October 2017; Published: 16 November 2017
Lens.org Logo
CITATIONS
Cited by 1 scholarly publication.
Advertisement
Advertisement
KEYWORDS
Luminescence

Tissues

Brain

Refractive index

Microscopy

Tissue optics

Green fluorescent protein

RELATED CONTENT

A new versatile clearing method for brain imaging
Proceedings of SPIE (March 10 2015)
Fast volumetric mapping of human brain slices
Proceedings of SPIE (April 01 2020)

Back to Top