Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 1408 word(s) 1408 2021-02-24 03:26:25 |
2 Format correct Meta information modification 1408 2021-02-25 06:35:40 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Wu, S. Amyloid Oligomers. Encyclopedia. Available online: https://encyclopedia.pub/entry/7553 (accessed on 27 April 2024).
Wu S. Amyloid Oligomers. Encyclopedia. Available at: https://encyclopedia.pub/entry/7553. Accessed April 27, 2024.
Wu, Si. "Amyloid Oligomers" Encyclopedia, https://encyclopedia.pub/entry/7553 (accessed April 27, 2024).
Wu, S. (2021, February 24). Amyloid Oligomers. In Encyclopedia. https://encyclopedia.pub/entry/7553
Wu, Si. "Amyloid Oligomers." Encyclopedia. Web. 24 February, 2021.
Amyloid Oligomers
Edit

Amyloid oligomers are considered to be potential targets for the development of therapeutic strategies for a wide range of neurodegenerative diseases. However, due to the low-populated, transient, and heterogeneous nature of amyloid oligomers, they are hard to characterize by conventional bulk methods. The development of single molecule approaches provides a powerful toolkit for investigating these oligomeric intermediates as well as the complex process of amyloid aggregation at molecular resolution.

Amyloid Oligomers Single molecule fluorescence detection protein aggregation neurodegenerative disease

1. Introduction

Cellular proteostasis is one of the key problems in biology and has been explored for decades. Within the complex cellular environment, newly synthesized polypeptides may be prone to misfold under certain conditions. Failing to eliminate misfolded proteins results in their accumulation and self-assembly into a variety of aggregates, such as the β-rich fibrillar structures termed amyloid [1][2][3]. The misfolding and aberrant aggregation of proteins and peptides into amyloid fibrils is associated with many human diseases, including the most common neurodegenerative diseases Alzheimer’s and Parkinson’s as well as type 2 diabetes [4][5]. In addition, the ability to form amyloid is found to be a common or generic property of polypeptide molecules and amyloid structures execute special biological functions in a range of cellular processes in living organisms [6][7]. Amyloid aggregation is a complicated process involving the transition from soluble monomers to insoluble aggregates. Although great effort and progress has been made to reveal the structures and assembly mechanisms of the β-sheet-rich amyloid fibrils [8][9][10][11][12], much evidence indicates that the oligomeric intermediates formed during the amyloid fibril formation process rather than mature fibrils are toxic to cells and are the major pathogenic agent in neurodegenerative disease [13][14][15][16]. Moreover, these oligomeric species with a wide range of sizes and distinct conformations are also crucial intermediates of amyloid aggregation. Therefore amyloid oligomers have attracted great interest during recent years and are considered to be potential targets for the development of therapeutic strategies [17].

It is desirable to study the structure, quantity and properties of oligomers during the process of amyloid aggregation. However, due to the low-populated, metastable and transient nature of amyloid oligomers, only limited information can be obtained using conventional bulk techniques such as the thioflavin T (ThT) fluorescence assay, circular dichroism spectroscopy (CD), dynamic light scattering (DLS), electron microscopy (EM) and nuclear magnetic resonance (NMR) spectroscopy, given that the oligomeric species are vastly outnumbered by monomeric and larger aggregate populations [18][19][20][21][22][23]. Fluorescence microscopy-based single molecule techniques have advantages over conventional biochemical and biophysical methods and can characterize individual protein molecules, allowing exploration of the highly heterogeneous and dynamic amyloid oligomers. By obtaining information at the single molecule level, species that represent only a small proportion of the system can nevertheless be detected, instead of being hidden and averaged as in the case of ensemble experiments, thus providing insight into the process of amyloid aggregation in unprecedented detail.

There are two primary experimental configurations for single molecule fluorescence detection, namely confocal microscopy and total internal reflection fluorescence (TIRF) microscopy. In the confocal setup, a collimated laser beam is focused by a high numerical aperture objective onto a diffraction-limited femtoliter volume. Fluorescence-labeled molecules at picomolar concentration in solution can then be excited to give fluorescence bursts when diffusing across the focal volume. In the TIRF setup, the laser is reflected by the bottom of the coverslip and the evanescent wave excites the fluorescence-labeled molecules that are immobilized or adsorbed on the surface of the coverslip. In contrast to the short observation time in the solution confocal setup (<1 ms), the surface-immobilized fluorescent molecules can be observed for longer time periods (seconds to minutes) before photobleaching. Under both setups, single-molecule Förster resonance energy transfer (smFRET) can be performed, in which a protein of interest is site-specifically labeled with a donor-acceptor dye pair and the distance change between the labeling sites is represented by the FRET efficiency. For most commercial dyes, FRET efficiency is sensitively dependent on the donor-acceptor distance in the range of 2–10 nm, thus is particularly suitable for investigating protein folding, structural transitions, and the populations and dynamics of specific protein conformations [24][25]. The confocal setup can also be employed for fluorescence correlation spectroscopy (FCS) which is based on the fluorescence fluctuations originating from the fluorescent molecules diffusing across the focal volume [26][27]. By analyzing the auto-correlation curves of fluorescence fluctuation, quantitative information such as the concentration, diffusional coefficient and dynamic properties of molecules can be obtained [28].

2. Characterization of Amyloid Oligomers by Single Molecule Fluorescence Detection

In order to detect the oligomeric species formed during amyloid fibrillization by smFRET, monomers are labeled with donor and acceptor fluorophores, mixed in equal amounts and then incubated to initiate aggregation. The formation of oligomers gives FRET signals that can be detected as coincident bursts in donor and acceptor detection channels when the oligomers diffuse across the focus even in the presence of an excess of monomers [29][30][31][32]. Two-color coincidence detection (TCCD) is an alternative detection mode, which similarly to smFRET is based on a confocal setup but is particularly suitable for samples with low intermolecular FRET efficiency. For TCCD the protein molecules are labeled with two spectrally separated fluorophores that are excited by an overlapped dual-color laser beam [33]. The fluorescence bursts occurring in only one emission channel are counted as monomers, while simultaneous fluorescence emission in both channels is counted as oligomers [34]. The oligomer concentration can be determined by converting the oligomer proportion referenced to a standard sample with known concentration. It is also possible to obtain information about the size and conformation of oligomers by analyzing the fluorescence intensity and FRET efficiency of oligomer bursts. During single molecule experiments, samples need to be diluted to the picomolar concentration range which may cause dissociation of unstable oligomeric species. In order to avoid this, fast-flow microfluidic techniques can be combined with single molecule detection, by which the acquisition time can be shortened from hours to minutes and so reduce the opportunity for oligomers to dissociate after dilution [35][36].

Amyloid oligomers can also be studied by TIRF imaging using either a single-color laser or coupled with TCCD. The samples are deposited onto the coverslip and excited by single-color or two overlapped lasers of different wavelengths. The intensity and colocalization of fluorescence in the images can be analyzed to give detailed information about the composition, size and morphology of the oligomeric species present [37]. An alternative to covalently labeling the target proteins with organic fluorophores, is to directly observe the unlabeled amyloid aggregates by TIRF imaging in the presence of a structure-specific dye such as ThT or pentameric formyl thiophene acetic acid (pFTAA) [38][39][40][41]. This can avoid exogenous effects introduced by covalently-linked dyes and photobleaching in conventional imaging. However, the resolution of traditional far-field optical imaging is around 250 nm due to the optical diffraction limit, far beyond the scale of amyloid oligomers and aggregates which are usually in the range from a few nanometers up to tens of nanometers. The development of super-resolution imaging techniques, either based on stochastic optical reconstruction microscopy (STORM) or stimulated emission depletion microscopy (STED), has now greatly improved the optical resolution to below 20 nm, allowing the exploration of the morphology and properties of amyloid aggregates on the nanoscale [42][43][44][45][46].

Over the past decade, with the development of ultrasensitive single molecule fluorescence techniques, amyloid oligomers have been thoroughly and quantitatively characterized, despite their low-populated, dynamic and heterogenous nature. The thermodynamic and kinetic properties of amyloid oligomers measured in vitro can be extrapolated to physiological conditions that are not accessible experimentally. This progress has greatly improved our understanding of the mechanism of amyloid aggregation and its relationship with neurodegenerative diseases. To date, the majority of single molecule studies on amyloid aggregation have been performed in vitro, but the simplified experimental conditions in vitro may not fully represent the more complicated cellular context. Due to the difficulty in achieving fluorescence labeling as well as improving the signal to noise ratio and fluorescence stability in vivo, it remains challenging to apply single molecule detection to living cells. However, a promising approach is the introduction of dye labels into living cells with the recent applications of microinjection or electroporation [47][48]. Meanwhile, advanced illumination techniques such as light-sheet or single-plane illumination microscopy that enable 3D sectioning with highly reduced background and limited phototoxicity provide opportunities for in vivo single molecule measurements of protein dynamics [49]. These and anticipated further development of new techniques will no doubt facilitate in vivo studies of the assembly of amyloid proteins at the single molecule level, therefore providing guidance on potential therapies for neurodegenerative diseases in the future.

References

  1. Ke, P.C.; Zhou, R.; Serpell, L.C.; Riek, R.; Knowles, T.P.J.; Lashuel, H.A.; Gazit, E.; Hamley, I.W.; Davis, T.P.; Fandrich, M.; et al. Half a century of amyloids: Past, present and future. Chem. Soc. Rev. 2020, 49, 5473–5509.
  2. Hartl, F.U. Protein misfolding diseases. Annu Rev. Biochem. 2017, 86, 21–26.
  3. Soto, C.; Pritzkow, S. Protein misfolding, aggregation, and conformational strains in neurodegenerative diseases. Nat. Neurosci. 2018, 21, 1332–1340.
  4. Dobson, C.M.; Knowles, T.P.J.; Vendruscolo, M. The amyloid phenomenon and its significance in biology and medicine. Cold Spring Harb. Perspect. Biol. 2020, 12, a033878.
  5. Chiti, F.; Dobson, C.M. Protein misfolding, amyloid formation, and human disease: A summary of progress over the last decade. Annu. Rev. Biochem. 2017, 86, 27–68.
  6. Pham, C.L.; Kwan, A.H.; Sunde, M. Functional amyloid: Widespread in Nature, diverse in purpose. Essays Biochem. 2014, 56, 207–219.
  7. Otzen, D.; Riek, R. Functional amyloids. Cold Spring Harb. Perspect. Biol. 2019, 11, a033860.
  8. Fitzpatrick, A.W.P.; Falcon, B.; He, S.; Murzin, A.G.; Murshudov, G.; Garringer, H.J.; Crowther, R.A.; Ghetti, B.; Goedert, M.; Scheres, S.H.W. Cryo-EM structures of tau filaments from Alzheimer’s disease. Nature 2017, 547, 185–190.
  9. Falcon, B.; Zhang, W.; Murzin, A.G.; Murshudov, G.; Garringer, H.J.; Vidal, R.; Crowther, R.A.; Ghetti, B.; Scheres, S.H.W.; Goedert, M. Structures of filaments from Pick’s disease reveal a novel tau protein fold. Nature 2018, 561, 137–140.
  10. Gremer, L.; Scholzel, D.; Schenk, C.; Reinartz, E.; Labahn, J.; Ravelli, R.B.G.; Tusche, M.; Lopez-Iglesias, C.; Hoyer, W.; Heise, H.; et al. Fibril structure of amyloid-β(1–42) by cryo-electron microscopy. Science 2017, 358, 116–119.
  11. Kollmer, M.; Close, W.; Funk, L.; Rasmussen, J.; Bsoul, A.; Schierhorn, A.; Schmidt, M.; Sigurdson, C.J.; Jucker, M.; Fandrich, M. Cryo-EM structure and polymorphism of Aβ amyloid fibrils purified from Alzheimer’s brain tissue. Nat. Commun. 2019, 10, 4760.
  12. Li, B.; Ge, P.; Murray, K.A.; Sheth, P.; Zhang, M.; Nair, G.; Sawaya, M.R.; Shin, W.S.; Boyer, D.R.; Ye, S.; et al. Cryo-EM of full-length α-synuclein reveals fibril polymorphs with a common structural kernel. Nat. Commun. 2018, 9, 3609.
  13. Cleary, J.P.; Walsh, D.M.; Hofmeister, J.J.; Shankar, G.M.; Kuskowski, M.A.; Selkoe, D.J.; Ashe, K.H. Natural oligomers of the amyloid-β protein specifically disrupt cognitive function. Nat. Neurosci. 2005, 8, 79–84.
  14. Winner, B.; Jappelli, R.; Maji, S.K.; Desplats, P.A.; Boyer, L.; Aigner, S.; Hetzer, C.; Loher, T.; Vilar, M.; Campioni, S.; et al. In vivo demonstration that α-synuclein oligomers are toxic. Proc. Natl. Acad. Sci. USA 2011, 108, 4194–4199.
  15. Bemporad, F.; Chiti, F. Protein misfolded oligomers: Experimental approaches, mechanism of formation, and structure-toxicity relationships. Chem. Biol. 2012, 19, 315–327.
  16. Conway, K.A.; Lee, S.J.; Rochet, J.C.; Ding, T.T.; Williamson, R.E.; Lansbury, P.T., Jr. Acceleration of oligomerization, not fibrillization, is a shared property of both α-synuclein mutations linked to early-onset Parkinson’s disease: Implications for pathogenesis and therapy. Proc. Natl. Acad. Sci. USA 2000, 97, 571–576.
  17. Guerrero-Munoz, M.J.; Castillo-Carranza, D.L.; Kayed, R. Therapeutic approaches against common structural features of toxic oligomers shared by multiple amyloidogenic proteins. Biochem. Pharmacol. 2014, 88, 468–478.
  18. LeVine, H., 3rd. Quantification of β-sheet amyloid fibril structures with thioflavin T. Methods Enzymol. 1999, 309, 274–284.
  19. Bartolini, M.; Bertucci, C.; Bolognesi, M.L.; Cavalli, A.; Melchiorre, C.; Andrisano, V. Insight into the kinetic of amyloid β (1–42) peptide self-aggregation: Elucidation of inhibitors’ mechanism of action. Chembiochem 2007, 8, 2152–2161.
  20. Newcomb, C.J.; Moyer, T.J.; Lee, S.S.; Stupp, S.I. Advances in cryogenic transmission electron microscopy for the characterization of dynamic self-assembling nanostructures. Curr. Opin. Colloid Interface Sci. 2012, 17, 350–359.
  21. Smith, J.F.; Knowles, T.P.; Dobson, C.M.; Macphee, C.E.; Welland, M.E. Characterization of the nanoscale properties of individual amyloid fibrils. Proc. Natl. Acad. Sci. USA 2006, 103, 15806–15811.
  22. Carulla, N.; Zhou, M.; Arimon, M.; Gairí, M.; Giralt, E.; Robinson, C.V.; Dobson, C.M. Experimental characterization of disordered and ordered aggregates populated during the process of amyloid fibril formation. Proc. Natl. Acad. Sci. USA 2009, 106, 7828–7833.
  23. Li, Y.; Lubchenko, V.; Vekilov, P.G. The use of dynamic light scattering and brownian microscopy to characterize protein aggregation. Rev. Sci. Instrum. 2011, 82, 053106.
  24. Lerner, E.; Cordes, T.; Ingargiola, A.; Alhadid, Y.; Chung, S.; Michalet, X.; Weiss, S. Toward dynamic structural biology: Two decades of single-molecule Forster resonance energy transfer. Science 2018, 359, eaan1133.
  25. Voith von Voithenberg, L.; Lamb, D.C. Single pair Forster resonance energy transfer: A versatile tool to investigate protein conformational dynamics. Bioessays 2018, 40, 1700078.
  26. Bacia, K.; Haustein, E.; Schwille, P. Fluorescence correlation spectroscopy: Principles and applications. Cold Spring Harb. Protoc. 2014, 2014, 709–725.
  27. Elson, E.L. Introduction to fluorescence correlation spectroscopy-brief and simple. Methods, 2018; 140–141, 3–9.
  28. Kitamura, A.; Kinjo, M. State-of-the-art fluorescence fluctuation-based spectroscopic techniques for the study of protein aggregation. Int. J. Mol. Sci. 2018, 19, 964.
  29. Cremades, N.; Cohen, S.I.; Deas, E.; Abramov, A.Y.; Chen, A.Y.; Orte, A.; Sandal, M.; Clarke, R.W.; Dunne, P.; Aprile, F.A.; et al. Direct observation of the interconversion of normal and toxic forms of α-synuclein. Cell 2012, 149, 1048–1059.
  30. Shammas, S.L.; Garcia, G.A.; Kumar, S.; Kjaergaard, M.; Horrocks, M.H.; Shivji, N.; Mandelkow, E.; Knowles, T.P.; Mandelkow, E.; Klenerman, D. A mechanistic model of tau amyloid aggregation based on direct observation of oligomers. Nat. Commun. 2015, 6, 7025.
  31. Iljina, M.; Garcia, G.A.; Horrocks, M.H.; Tosatto, L.; Choi, M.L.; Ganzinger, K.A.; Abramov, A.Y.; Gandhi, S.; Wood, N.W.; Cremades, N.; et al. Kinetic model of the aggregation of α-synuclein provides insights into prion-like spreading. Proc. Natl. Acad. Sci. USA 2016, 113, E1206–E1215.
  32. Yang, J.; Dear, A.J.; Michaels, T.C.T.; Dobson, C.M.; Knowles, T.P.J.; Wu, S.; Perrett, S. Direct observation of oligomerization by single molecule fluorescence reveals a multistep aggregation mechanism for the yeast prion protein Ure2. J. Am. Chem. Soc. 2018, 140, 2493–2503.
  33. Orte, A.; Clarke, R.; Klenerman, D. Single-molecule two-colour coincidence detection to probe biomolecular associations. Biochem. Soc. Trans. 2010, 38, 914–918.
  34. Orte, A.; Birkett, N.R.; Clarke, R.W.; Devlin, G.L.; Dobson, C.M.; Klenerman, D. Direct characterization of amyloidogenic oligomers by single-molecule fluorescence. Proc. Natl. Acad. Sci. USA 2008, 105, 14424–14429.
  35. Horrocks, M.H.; Li, H.; Shim, J.-U.U.; Ranasinghe, R.T.; Clarke, R.W.; Huck, W.T.; Abell, C.; Klenerman, D. Single molecule fluorescence under conditions of fast flow. Anal. Chem. 2012, 84, 179–185.
  36. Horrocks, M.H.; Tosatto, L.; Dear, A.J.; Garcia, G.A.; Iljina, M.; Cremades, N.; Dalla Serra, M.; Knowles, T.P.; Dobson, C.M.; Klenerman, D. Fast flow microfluidics and single-molecule fluorescence for the rapid characterization of α-synuclein oligomers. Anal. Chem. 2015, 87, 8818–8826.
  37. Narayan, P.; Orte, A.; Clarke, R.W.; Bolognesi, B.; Hook, S.; Ganzinger, K.A.; Meehan, S.; Wilson, M.R.; Dobson, C.M.; Klenerman, D. The extracellular chaperone clusterin sequesters oligomeric forms of the amyloid-β(1-40) peptide. Nat. Struct. Mol. Biol. 2012, 19, 79–83.
  38. Varela, J.A.; Rodrigues, M.; De, S.; Flagmeier, P.; Gandhi, S.; Dobson, C.M.; Klenerman, D.; Lee, S.F. Optical structural analysis of individual α-synuclein oligomers. Angew. Chem. Int. Ed. Engl. 2018, 57, 4886–4890.
  39. Horrocks, M.H.; Lee, S.F.; Gandhi, S.; Magdalinou, N.K.; Chen, S.W.; Devine, M.J.; Tosatto, L.; Kjaergaard, M.; Beckwith, J.S.; Zetterberg, H.; et al. Single-Molecule Imaging of Individual Amyloid Protein Aggregates in Human Biofluids. ACS Chem. Neurosci. 2016, 7, 399–406.
  40. Sang, J.C.; Meisl, G.; Thackray, A.M.; Hong, L.; Ponjavic, A.; Knowles, T.P.J.; Bujdoso, R.; Klenerman, D. Direct observation of murine prion protein replication in vitro. J. Am. Chem. Soc. 2018, 140, 14789–14798.
  41. Ries, J.; Udayar, V.; Soragni, A.; Hornemann, S.; Nilsson, K.P.; Riek, R.; Hock, C.; Ewers, H.; Aguzzi, A.A.; Rajendran, L. Superresolution imaging of amyloid fibrils with binding-activated probes. ACS Chem. Neurosci. 2013, 4, 1057–1061.
  42. Bongiovanni, M.N.; Godet, J.; Horrocks, M.H.; Tosatto, L.; Carr, A.R.; Wirthensohn, D.C.; Ranasinghe, R.T.; Lee, J.E.; Ponjavic, A.; Fritz, J.V.; et al. Multi-dimensional super-resolution imaging enables surface hydrophobicity mapping. Nat. Commun. 2016, 7, 13544.
  43. Lee, J.E.; Sang, J.C.; Rodrigues, M.; Carr, A.R.; Horrocks, M.H.; De, S.; Bongiovanni, M.N.; Flagmeier, P.; Dobson, C.M.; Wales, D.J.; et al. Mapping surface hydrophobicity of α-synuclein oligomers at the nanoscale. Nano Lett. 2018, 18, 7494–7501.
  44. Sang, J.C.; Lee, J.E.; Dear, A.J.; De, S.; Meisl, G.; Thackray, A.M.; Bujdoso, R.; Knowles, T.P.J.; Klenerman, D. Direct observation of prion protein oligomer formation reveals an aggregation mechanism with multiple conformationally distinct species. Chem. Sci. 2019, 10, 4588–4597.
  45. Spehar, K.; Ding, T.; Sun, Y.; Kedia, N.; Lu, J.; Nahass, G.R.; Lew, M.D.; Bieschke, J. Super-resolution imaging of amyloid structures over extended times by using transient binding of single thioflavin T molecules. Chembiochem 2018, 19, 1944–1948.
  46. Torra, J.; Bondia, P.; Gutierrez-Erlandsson, S.; Sot, B.; Flors, C. Long-term STED imaging of amyloid fibers with exchangeable thioflavin T. Nanoscale 2020, 12, 15050–15053.
  47. König, I.; Zarrine-Afsar, A.; Aznauryan, M.; Soranno, A.; Wunderlich, B.; Dingfelder, F.; Stüber, J.C.; Plückthun, A.; Nettels, D.; Schuler, B. Single-molecule spectroscopy of protein conformational dynamics in live eukaryotic cells. Nat. Methods 2015, 12, 773–779.
  48. Crawford, R.; Torella, J.P.; Aigrain, L.; Plochowietz, A.; Gryte, K.; Uphoff, S.; Kapanidis, A.N. Long-lived intracellular single-molecule fluorescence using electroporated molecules. Biophys. J. 2013, 105, 2439–2450.
  49. Aigrain, L.; Sustarsic, M.; Crawford, R.; Plochowietz, A.; Kapanidis, A.N. Internalization and observation of fluorescent biomolecules in living microorganisms via electroporation. J. Vis. Exp. 2015, 52208.
More
Information
Subjects: Biophysics
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 391
Revisions: 2 times (View History)
Update Date: 21 Apr 2021
1000/1000