Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-8bljj Total loading time: 0 Render date: 2024-06-22T13:10:07.417Z Has data issue: false hasContentIssue false

6 - Posttranslational Protein Translocation through Membranes at the Single-Molecule Level

from Part II - Protein Folding, Structure, Confirmation, and Dynamics

Published online by Cambridge University Press:  05 May 2022

Krishnarao Appasani
Affiliation:
GeneExpression Systems, Inc.
Raghu Kiran Appasani
Affiliation:
Psychiatrist, Neuroscientist, & Mental Health Advocate
Get access

Summary

Protein secretion studies started in the 1950s with George Palade’s electron microscopy (EM) work (Palade, 1952, 1975). Protein secretion is a very relevant process because more than 30 percent of synthesized proteins work in organelles or outside the cells (Arora and Tamm, 2001). In eukaryotic cells, the proteins secreted to the exterior are synthesized in the cytoplasm and transported inside the endoplasmic reticulum (ER), then pass to the Golgi apparatus and finally to secretory vesicles. Blobel and Sabatini in the 1970s discovered signal sequences at the N-terminus extreme of secretory proteins that allow them to be recognized by receptors thus mediating and facilitating their entrance to ER interior (Blobel and Dobberstein, 1975; Sabatini et al., 1982). Proteins enter the ER lumen by a protein conducting channel formed by a protein complex, known as the translocon, discovered in yeast in Randy Schekman´s laboratory, which is universally conserved (Deshaies et al., 1991). In eukaryotic cells, the translocation of proteins into ER lumen is carried out by the Sec61 complex (Rapoport, 2007; Zimmermann et al., 2011), whereas the bacterial homologue is the heterotrimeric SecY complex, which allows the secretion of proteins to the exterior (Park and Rapoport, 2012).

Type
Chapter
Information
Single-Molecule Science
From Super-Resolution Microscopy to DNA Mapping and Diagnostics
, pp. 80 - 94
Publisher: Cambridge University Press
Print publication year: 2022

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Arora, A. and Tamm, L. K. (2001). Byophisical Approaches to Membrane Protein Structure Determination. Current Opinion in Structural Biology, 11, 540547.Google Scholar
a Nijeholt, J. A. L. and Driessen, A. J. (2012). The Bacterial Sec-Translocase: Structure and Mechanism. Philosophical Transactions of the Royal Society B, 367, 10161028.Google Scholar
Ashkin, A. (1970). Acceleration and Trapping of Particles by Radiation Pressure. Physical Review Letters, 24, 156159.CrossRefGoogle Scholar
Ashkin, A., Dziedzic, J. M., Bjorkholm, J. E., et al. (1986). Observation of a Single-Beam Gradient Force Optical Trap for Dielectric Particles. Optics Letters, 11, 288290.Google Scholar
Astumian, R. D. (1997). Thermodynamics and Kinetics of a Brownian Motor. Science, 276, 917922.Google Scholar
Banerjee, R., Jayaraj, G. G., Peter, J. J., et al. (2016). Monitoring Conformational Heterogeneity of the Lid of DnaK Substrate-Binding Domain during Chaperone Cycle. FEBS Journal, 283, 28532868.Google Scholar
Bechtluft, P, Van Leeuwen, R. G., Tyreman, M., et al. (2007). Direct Observation of Chaperone-Induced Changes in a Protein Folding Pathway. Science, 318, 14581461.Google Scholar
Becker, T., Bhushan, S., Jarasch, A., et al. (2009). Structure of Monomeric Yeast and Mammalian Sec61 Complexes Interacting with the Translating Ribosome. Science, 326, 13691373.Google Scholar
Behnke, J., Feige, M. J., and Hendershot, L. M. (2015). BiP and Its Nucleotide Exchange Factors Grp170 and Sil1: Mechanisms of Action and Biological Functions. Journal of Molecular Biology, 427, 15891608.Google Scholar
Bertelsen, E. B., Chang, L., Gestwicki, J. E., et al. (2009). Solution Conformation of Wild-Type E. coli Hsp70 (DnaK) Chaperone Complexed with ADP and Substrate. Proceedings of the National Acadamy of Science U.S.A., 106, 84718476.Google Scholar
Bertz, M. and Rief, M. (2009). Ligand Binding Mechanics of Maltose Binding Protein. Journal of Molecular Biology, 393, 10971105.Google Scholar
Blobel, G. and Dobberstein, B. (1975). Transfer of Proteins across Membranes. I. Presence of Proteolytically Processed and Unprocessed Nascent Immunoglobulin Light-Chains on Membrane-Bound Ribosomes of Murine Myeloma. Journal of Cell Biology, 67, 835851.CrossRefGoogle ScholarPubMed
Bustamante, C. (2008). In Singulo Biochemistry: When Less Is More. Annual Review of Biochemistry, 77, 4550.CrossRefGoogle ScholarPubMed
Bustamante, C., Chemla, Y. R., Forde, N. R., et al. (2004). Mechanical Processes in Biochemistry. Annual Review of Biochemistry, 73, 705748.Google Scholar
Bustamante, C., Cheng, W., and Mejia, Y. X. (2011). Revisiting the Central Dogma One Molecule at a Time. Cell, 144, 480497.Google Scholar
Bustamante, C., Kaiser, C. M., Maillard, R. A., et al. (2014). Mechanisms of Cellular Proteostasis: Insights from Single-Molecule Approaches. Annual Review of Biophysics, 43, 119140.Google Scholar
Bustamante, C., Macosko, J. C., and Wuite, G. J. (2000). Grabbing the Cat by the Tail: Manipulating Molecules One by One. Nature Reviews Molecular Cell Biology, 1, 130136.Google Scholar
Cecconi, C., Shank, E. A., Bustamante, C., et al. (2005). Direct Observation of the Three-State Folding of a Single Molecule. Science, 309, 20572060.Google Scholar
Cecconi, C., Shank, E. A., Marqusee, S., et al. (2007). Studying Protein Folding with Laser Tweezers. Proceedings of the International School of Physics “Enrico Fermi,” 165, 145160.Google Scholar
Cecconi, C., Shank, E. A., Dahlquist, F. W., et al. (2008). Protein-DNA Chimeras for Single Molecule Mechanical Folding Studies with the Optical Tweezers. European Biophysics Journal, 37, 729733.Google Scholar
del Rio, A., Perez-Jimenez, R., Liu, R., et al. (2009). Stretching Single Talin Rod Molecules Activates Vinculin Binding. Science, 323, 638641.Google Scholar
Deniz, A. A., Mukhopadhyay, S., and Lemke, E. A. (2008). Single-Molecule Biophysics: At the Interface of Biology, Physics and Chemistry. Journal of the Royal Society Interface, 5, 1545.Google Scholar
Deshaies, R. J., Sanders, S. L., Feldheim, D. A., et al. (1991). Assembly of Yeast Sec Proteins Involved in Translocation into the Endoplasmic Reticulum into a Membrane-Bound Multisubunit Complex. Nature, 349, 806808.Google Scholar
Erlandson, K. J., Millar, S. B. M., Nam, Y., et al. (2008). A Role for the Two-Helix Finger of the SecA ATPase in Protein Translocation. Nature, 455, 984987.Google Scholar
Fisher, T. E., Marszalek, P. E., and Fernandez, J. M. (2000). Stretching Single Molecules into Novel Conformations Using the Atomic Force Microscope. Nature Structural and Molecular Biology, 7, 719724.Google Scholar
Frauenfeld, J., Gumbart, J., Sluis, E. O., et al. (2011). Cryo-EM Structure of the Ribosome-SecYE Complex in the Membrane Environment. Nature Structural and Molecular Biology, 18, 614621.Google Scholar
Gogala, M., Becker, T., Beatrix, B., et al. (2014). Structures of the Sec61 Complex Engaged in Nascent Peptide Translocation or Membrane Insertion. Nature, 506, 107110.Google Scholar
Goloubinoff, P. and De los Ríos, P. (2007). The Mechanism of Hsp70 Chaperones: (Entropic) Pulling the Models Together. Trends in Biochemical Science, 32, 372380.Google Scholar
Guo, Q., He, Y., and Lu, H. P. (2015). Interrogating the Activities of Conformational Deformed Enzyme by Single-Molecule Fluorescence-Magnetic Tweezers Microscopy. Proceedings of the National Academy of Science U.S.A., 112, 1390413909.Google Scholar
Junker, J. P., Hell, K., Schlierf, M., et al. (2005). Influence of Substrate Binding on the Mechanical Stability of Mouse Dihydrofolate Reductase. Biophysics Journal, 89, L46L48.Google Scholar
Junker, J. P., Ziegler, F., and Rief, M. (2009). Ligand-Dependent Equilibrium Fluctuations of Single Calmodulin Molecules. Science, 323, 633637.Google Scholar
Kainov, D. E., Tuma, R., and Mancini, E. J. (2006). Hexameric Molecular Motors: P4 Packaging ATPase Unravels the Mechanism. Cellular and Molecular Life Sciences, 63, 10951105.Google Scholar
Kedrov, A., Kusters, I., Krasnikov, V. V., et al. (2011). A Single Copy of SecYEG Is Sufficient for Preprotein Translocation. EMBO Journal, 30, 43874397.CrossRefGoogle ScholarPubMed
Kellner, R., Hofmann, H., Barducci, A., et al. (2014). Single-Molecule Spectroscopy Reveals Chaperone-Mediated Expansion of Substrate Protein. Proceedings of the National Academy of Science U.S.A.,111, 1335513360.Google Scholar
Kosakowska-Cholody, T., Lin, J., Srideshikan, S. M., et al. (2014). HKH40A Downregulates GRP78/BiP Expression in Cancer Cells. Cell Death and Disease, 5, e1240.Google Scholar
Kusters, I., van den Bogaart, G., Kedrov, A., et al. (2011). Quaternary Structure of SecA in Solution and Bound to SecYEG Probed at the Single Molecule Level. Structure, 19 , 430439.CrossRefGoogle ScholarPubMed
Latorre, R., Ehrenstein, G., and Lecar, H. (1972). Ion Transport through Excitability-Inducing Material (EIM) Channels in Lipid Bilayer Membranes. Journal of General Physiology, 60, 7285.Google Scholar
Lee, A. S. (2014). Glucose-Regulated Proteins in Cancer: Molecular Mechanisms and Therapeutic Potential. Nature Reviews Cancer, 14, 263276.Google Scholar
Li, G. W. and Xie, X. S. (2011). Central Dogma at the Single-Molecule Level in Living Cells. Nature, 475, 308315.Google Scholar
Li, L., Park, E., Ling, J., et al. (2016). Crystal Structure of a Substrate-Engaged SecY Protein-Translocation Channel. Nature, 531, 395399.Google Scholar
Lyubimov, A. Y., Strycharska, M., and Berger, J. M. (2011). The Nuts and Bolts of Ring-Translocase Structure and Mechanism. Current Opinion in Structural Biology, 21, 240248.Google Scholar
Maillard, R. A., Chistol, G., Sen, M., et al. (2011). ClpX(P) Generates Mechanical Force to Unfold and Translocate Its Protein Substrates. Cell, 145 , 459469.Google Scholar
Mapa, K., Sikor, M., Kudryavtsev, V., et al. (2010). The Conformational Dynamics of the Mitochondrial Hsp70 Chaperone. Molecular Cell, 38, 89100.Google Scholar
Marcinowski, M., Höller, M., Feige, M. J., et al. (2011). Substrate Discrimination of the Chaperone BiP by Autonomous and Cochaperone-Regulated Conformational Transitions. Nature Structural and Molecular Biology, 18 , 150158.Google Scholar
Mashaghi, A., Bezrukavnikov, S., Minde, D. P., et al. (2016). Alternative Modes of Client Binding Enable Functional Plasticity of Hsp70. Nature, 539, 448451.Google Scholar
Matlack, K. E., Misselwitz, B., Plath, K., et al. (1999). BiP Acts as a Molecular Ratchet during Post-Translational Transport of Prepo-αfactor across the ER Membrane. Cell, 97, 553564.Google Scholar
Min, D., Jefferson, R. E., Bowie, J. U., et al. (2015). Mapping the Energy Landscape for Second-Stage Folding of a Single Membrane Protein. Nature Chemical Biology, 11, 981987.Google Scholar
Neher, E. and Sakmann, B., (1976). Single-Channel Currents Recorded from Membrane of Denervated Frog Muscle Fibres. Nature, 260, 799802.Google Scholar
Palade, G. (1952). A Study of Fixation for Electron Microscopy. Journal of Experimental Medicine, 95 , 285297.Google Scholar
Palade, G. (1975). Intracellular Aspects of the Process of Protein Synthesis. Science, 189, 347358.Google Scholar
Park, E. and Rapoport, T. A. (2012). Mechanism of Sec61/SecY-Mediated Protein Translocation across Membranes. Annual Review of Biophysics, 41, 120.Google Scholar
Ramírez, M. P., Rivera, M., Quiroga-Roger, D., et al. (2017). Single Molecule Force Spectroscopy Reveals the Effect of BiP Chaperone on Protein Folding. Protein Science, 26, 14041412.Google Scholar
Rapoport, T. A. (2007). Protein Translocation across the Eukaryotic Endoplasmic Reticulum and Bacterial Plasma Membranes. Nature, 450, 663669.Google Scholar
Rapoport, T. A., Li, L., and Park, E. (2017). Structural and Mechanistic Insights into Protein Translocation. Annual Review of Cell and Development. Biology, 33, 369390.Google Scholar
Sabatini, D. D., Kreibich, G., Morimoto, T., et al. (1982). Mechanism for the Incorporation of Proteins in Membranes and Organelles. Journal of Cell Biology, 92, 122.Google Scholar
Saparov, S. M., Erlandson, K., Cannon, K., et al. (2007). Determining the Conductance of the SecY Protein Translocation Channel for Small Molecules. Molecular Cell, 26, 501509.Google Scholar
Schekman, R., (1994). Translocation gets a push. Cell, 78, 911913.Google Scholar
Schwille, P., Meyer-Almes, F.J., Rigler, R., (1997). Dual-color fluorescence cross-correlation spectroscopy for multicomponent diffusional analysis in solution. Biophys J., 72, 18781886.Google Scholar
Shank, E. A., Cecconi, C., Dill, J. W., et al. (2010). The Folding Cooperativity of a Protein Is Controlled by Its Chain Topology. Nature, 465, 637640.Google Scholar
Shields, A. M., Panayi, G. S., and Corrigall, V. M., (2012). A New-Age for Biologic Therapies: Long-Term Drug-Free Therapy with BiP? Frontiers in Immunology, 3, 18.Google Scholar
Smith, D. E., Tans, S. J., Smith, S. B., et al. (2001). The Bacteriophage Straight phi29 Portal Motor Can Package DNA against a Large Internal Force. Nature, 413, 748752.Google Scholar
Taufik, I., Kedrov, A., Exterkate, M., et al. (2013). Monitoring the Activity of Single Translocons. Journal of Molecular Biology, 425, 41454153.Google Scholar
Tinoco, I. and Gonzalez, R. L. (2011). Biological Mechanisms, One Molecule at a Time. Genes and Development, 25, 12051231.Google Scholar
Tsai, Y. L., Zhang, Y., Tseng, C. C., et al. (2015). Characterization and Mechanism of Stress-Induced Translocation of 78-Kilodalton Glucose-Regulated Protein (GRP78) to the Cell Surface. Journal of Biological Chemistry, 290 , 80498064.Google Scholar
Wu, Z. C., de Keyzer, J., Kedrov, A., et al. (2012). Competitive Binding of the SecA ATPase and Ribosomes to the SecYEG Translocon. Journal of Biological Chemistry, 287, 78857895.Google Scholar
Yang, J., Nune, M., Zong, Y., et al. (2015). Close and Allosteric Opening of the Polypeptide-Binding Site in a Human Hsp70 Chaperone BiP. Structure, 23, 21912203.Google Scholar
Zhang, X., Halvorsen, K., Zhang, C. Z., et al. (2009). Mechanoenzymatic Cleavage of the Ultralarge Vascular Protein Von Willebrand Factor. Science, 324, 13301334.Google Scholar
Zhang, Y., Tseng, C. C., Tsai, Y. L., et al. (2013). Cancer Cells Resistant to Therapy Promote Cell Surface Relocalization of GRP78 Which Complexes with PI3K and Enhances PI(3,4,5)P3 Production. PLoS One, 8, e80071.Google Scholar
Zimmer, J., Nam, Y., and Rapoport, T. A. (2008). Structure of a Complex of the ATPase SecA and the Protein-Translocation Channel. Nature, 455, 936943.Google Scholar
Zimmermann, R., Eyrisch, S., Asmad, M., et al. (2011). Protein Translocation across the ER Membrane. Biochimica et Biophysica Acta, 1808, 912924.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×