Synaptotagmin

Last updated
Synaptotagmin
Exocytosis-machinery.jpg
Molecular machinery driving exocytosis in neurotransmitter release: the core SNARE complex (formed by four α-helices contributed by synaptobrevin, syntaxin and SNAP-25) and the Ca2+ sensor synaptotagmin. [1]
Identifiers
SymbolSYT
OPM superfamily 45
OPM protein 3hn8
Membranome 199

Synaptotagmins (SYTs) constitute a family of membrane-trafficking proteins that are characterized by an N-terminal transmembrane region (TMR), a variable linker, and two C-terminal C2 domains - C2A and C2B. There are 17 isoforms in the mammalian synaptotagmin family. [2] There are several C2-domain containing protein families that are related to synaptotagmins, including transmembrane (Ferlins, Extended-Synaptotagmin (E-Syt) membrane proteins, and MCTPs) and soluble (RIMS1 and RIMS2, UNC13D, synaptotagmin-related proteins and B/K) proteins. The family includes synaptotagmin 1, a Ca2+ sensor in the membrane of the pre-synaptic axon terminal, coded by gene SYT1. [3]

Contents

Functions

Based on their brain/endocrine distribution and biochemical properties, in particular C2 domains of certain synaptotagmins bound to calcium, synaptotagmins were proposed to function as calcium sensors in the regulation of neurotransmitter release and hormone secretion. Although synaptotagmins share a similar domain structure and a high degree of homology in the C2 domains, not all synaptotagmins bind to calcium. In fact, only eight out of the fifteen synaptotagmins are capable of calcium binding. The calcium binding synaptotagmins include synaptotagmins 1, 2, 3, 5, 6, 7, 9, and 10. The remaining seven synaptotagmins do not bind to calcium due to the lack of calcium coordinating residues or spatial orientation of the acidic residues (see the section on C2 domains for details).

Calcium-binding synaptotagmins act as Ca2+ sensors and are involved in both:

  1. early synaptic vesicle docking to the presynaptic membrane via interaction with β-neurexin [4] or SNAP-25 [5]
  2. late steps of Ca2+-evoked synaptic vesicle fusion with the presynaptic membrane. [6] [7] [8] It was also shown that synaptotagmin 1 can displace complexin from the SNARE complex in the presence of calcium. This is thought to be one of the last steps in exocytosis. [9] Calcium-bound synaptotagmin binding to the SNARE complex, causes the fusion clamp effect of complexin to be released, allowing vesicle fusion to occur and exocytosis to proceed. [10]

Synaptotagmins directly affect the synchronicity of calcium-dependent neurotransmission. While the suppression of Syt1 blocks fast, synchronous neurotransmission, it also enhances slow, asynchronous neurotransmission. [11] On the other hand, suppression of Syt7 hinders the slower, asynchronous release of neurotransmitters. This suggests that synaptotagmin-7 is responsible for mediating a slower form of Ca(2+)-triggered release while the faster release is induced by synaptotagmin-1. These discrepancies illustrate important distinctions between synaptotagmin isoforms and how they underlie the kinetics of neurotransmission and long-term potentiation.

C-terminal C2-domains

The C2 domain is a widely occurring conserved sequence motif of 130-140 amino acid residues, which was first defined as the second constant sequence in PKC isoforms. [12] The C2 domain was first shown to bind to calcium in synaptotagmin-1. Subsequent atomic structure analysis of synaptotagmin-1 at 1.9 Å resolution indicated that its C2 domains are composed of a stable eight-stranded β-sandwich with flexible loops emerging from the top and bottom. Nuclear magnetic resonance (NMR) studies of synaptotagmin-1 revealed that calcium binds exclusively to the top loops, and the binding pockets are coordinated by five conserved aspartate residues: three calcium ions bind to C2A via D172, D178, D230, D232, S235 and D238, and two calcium ions bind to C2B via D303, D309, D363, 365 and D371. Not all synaptotagmin C2 domains bind to calcium. In fact, based on sequence similarities and subsequent confirmation by biochemical analyses, only eight synaptotagmins bind to calcium, namely, synaptotagmin-1, -2, -3, -5, -6, -7, -9 and -10. The lack of critical residues involved in calcium binding accounts for the majority of failure in calcium-binding by the other synaptotagmins. This includes both C2 domains of synaptotagmin-11, -12, -13, -14 and -15, and C2A domain of synaptotagmin-4 and -8. Synaptotagmin-4 and -11 C2B domains, which possess all five acidic residues in the top loops, however, do not bind to calcium due to spatial orientation of the calcium ligands that fail to form proper calcium binding sites. For calcium-binding synaptotagmins, although amino acid residues in the top loops other than those mentioned above are not directly involved in coordinating calcium binding, they affect calcium binding affinity, such as R233 in synaptotagmin-1. The diversity of sequences and structures flanking the calcium-coordinating amino acid residues renders the eight synaptotagmins bind to calcium at various affinities, covering the full range of calcium requirements for regulated exocytosis.

The C2A domain regulates the fusion step of synaptic vesicle exocytosis. [13] [14] Consistent with this, the kinetics of Ca2+-dependent phospholipid binding activity of the C2A domain in vitro are compatible with the very fast nature of neurotransmitter release (within 200 μs). [15] The C2A domain was shown to bind negatively charged phospholipids in a Ca2+-dependent fashion. Ca2+-binding alters the protein-protein interactions of synaptotagmin such as increasing the affinity of synaptotagmin for syntaxin.

The C2B domain binds to phosphatidyl-inositol-3,4,5-triphosphate (PIP3) in the absence of calcium ions, and to phosphatidylinositol bisphosphate (PIP2) in their presence, [16] suggesting that a lipid-interaction switch occurs during depolarization. Ca2+-binding to the C2B domain confers synaptotagmin dimerization involved in the fusion step of synaptic vesicles by Ca2+-dependent self-clustering via the C2B domain. [17] Ca2+-independent is the interaction between the C2B domain and SNAP-25, and between the C2B domain and the "synprint" (synaptic protein interaction) motif of the pore-forming subunit of voltage-gated calcium channels. The C2B domain regulates also the recycling step of synaptic vesicles by binding to the clathrin assembly protein, AP-2.

Plasticity and Learning

Synaptotagmin variants have been implicated in the enhancement of neural connections, leading to long-term potentiation(LTP) in synapses. The localization of synaptotagmin to the endoplasmic reticulum in the cytoplasm drives the growth of these synapses. [18] Synaptogmins such as Syt1 and Syt7 also play a role in calcium-dependent AMPA receptors exocytosis to the neuron membrane. [19] This process initiates LTP formation and underlies learning. Moreover, synaptotagmins are able to respond to elevated levels of calcium at synapses during single action potentials by further heightening calcium levels via withdrawal from intracellular stores. [18] This leads to a stronger response in the postsynaptic cell.

Other important roles

Synaptotagmins have been shown to regulate exocytosis in other intracellular organelles such as lysosomes. [20] Suppression of Syt7 in astrocytes prevents injury repairment through weakened lysosome exocytosis, suggesting a role of synaptotagmin proteins in mediating repair following brain damage by interacting with lysosomes.

Apart from the molecular events mediated by synaptotagmins, these proteins have also been identified to play a large role in the cognitive realm. Bipolar disorder is an example of one such instance where synaptotagmins exhibit their effects in a larger context. Knockout of synaptotagmin proteins in animal models elicited both manic and depressive-like symptoms, characteristic of BD. [21] This suggests that synaptotagmin depletion is associated with BD pathology.

Members

References and notes

  1. Georgiev DD, Glazebrook JF (2007). "Subneuronal processing of information by solitary waves and stochastic processes". In Lyshevski SE (ed.). Nano and Molecular Electronics Handbook. Nano and Microengineering Series. CRC Press. pp. 17–1–17-41. doi:10.1201/9781315221670-17. ISBN   978-0-8493-8528-5.
  2. Dean C, Dunning FM, Liu H, Bomba-Warczak E, Martens H, Bharat V, et al. (May 2012). "Axonal and dendritic synaptotagmin isoforms revealed by a pHluorin-syt functional screen". Molecular Biology of the Cell. 23 (9): 1715–27. doi:10.1091/mbc.E11-08-0707. PMC   3338438 . PMID   22398727.
  3. "NIH VideoCast - Ca2+ Sensors for Exocytosis". videocast.nih.gov. Retrieved 16 April 2018.
  4. Fukuda M, Moreira JE, Liu V, Sugimori M, Mikoshiba K, Llinás RR (December 2000). "Role of the conserved WHXL motif in the C terminus of synaptotagmin in synaptic vesicle docking". Proceedings of the National Academy of Sciences of the United States of America. 97 (26): 14715–9. Bibcode:2000PNAS...9714715F. doi: 10.1073/pnas.260491197 . PMC   18984 . PMID   11114192.
  5. Schiavo G, Stenbeck G, Rothman JE, Söllner TH (February 1997). "Binding of the synaptic vesicle v-SNARE, synaptotagmin, to the plasma membrane t-SNARE, SNAP-25, can explain docked vesicles at neurotoxin-treated synapses". Proceedings of the National Academy of Sciences of the United States of America. 94 (3): 997–1001. Bibcode:1997PNAS...94..997S. doi: 10.1073/pnas.94.3.997 . PMC   19628 . PMID   9023371.
  6. Pang ZP, Melicoff E, Padgett D, Liu Y, Teich AF, Dickey BF, et al. (December 2006). "Synaptotagmin-2 is essential for survival and contributes to Ca2+ triggering of neurotransmitter release in central and neuromuscular synapses". The Journal of Neuroscience. 26 (52): 13493–504. doi:10.1523/JNEUROSCI.3519-06.2006. PMC   6674714 . PMID   17192432.
  7. Maximov A, Südhof TC (November 2005). "Autonomous function of synaptotagmin 1 in triggering synchronous release independent of asynchronous release". Neuron. 48 (4): 547–54. doi: 10.1016/j.neuron.2005.09.006 . PMID   16301172.
  8. O'Connor V, Lee AG (September 2002). "Synaptic vesicle fusion and synaptotagmin: 2B or not 2B?". Nature Neuroscience. 5 (9): 823–4. doi:10.1038/nn0902-823. PMID   12196805. S2CID   39531338.
  9. Tang J, Maximov A, Shin OH, Dai H, Rizo J, Südhof TC (September 2006). "A complexin/synaptotagmin 1 switch controls fast synaptic vesicle exocytosis". Cell. 126 (6): 1175–87. doi: 10.1016/j.cell.2006.08.030 . PMID   16990140.
  10. Maximov A, Tang J, Yang X, Pang ZP, Südhof TC (January 2009). "Complexin controls the force transfer from SNARE complexes to membranes in fusion". Science. 323 (5913): 516–21. Bibcode:2009Sci...323..516M. doi:10.1126/science.1166505. PMC   3235366 . PMID   19164751.
  11. Bacaj, Taulant; Wu, Dick; Yang, Xiaofei; Morishita, Wade; Zhou, Peng; Xu, Wei; Malenka, Robert C.; Südhof, Thomas C. (2013-11-20). "Synaptotagmin-1 and synaptotagmin-7 trigger synchronous and asynchronous phases of neurotransmitter release". Neuron. 80 (4): 947–959. doi:10.1016/j.neuron.2013.10.026. ISSN   1097-4199. PMC   3888870 . PMID   24267651.
  12. Kohout, Susy C.; Corbalán-García, Senena; Torrecillas, Alejandro; Goméz-Fernandéz, Juan C.; Falke, Joseph J. (2002-09-24). "C2 domains of protein kinase C isoforms alpha, beta, and gamma: activation parameters and calcium stoichiometries of the membrane-bound state". Biochemistry. 41 (38): 11411–11424. doi:10.1021/bi026041k. ISSN   0006-2960. PMC   3640336 . PMID   12234184.
  13. Zimmerberg J, Akimov SA, Frolov V (April 2006). "Synaptotagmin: fusogenic role for calcium sensor?". Nature Structural & Molecular Biology. 13 (4): 301–3. doi:10.1038/nsmb0406-301. PMID   16715046. S2CID   32067429.
  14. Fernández-Chacón R, Königstorfer A, Gerber SH, García J, Matos MF, Stevens CF, et al. (March 2001). "Synaptotagmin I functions as a calcium regulator of release probability". Nature. 410 (6824): 41–9. Bibcode:2001Natur.410...41F. doi:10.1038/35065004. PMID   11242035. S2CID   1756258.
  15. Chapman ER (July 2002). "Synaptotagmin: a Ca(2+) sensor that triggers exocytosis?". Nature Reviews. Molecular Cell Biology. 3 (7): 498–508. doi:10.1038/nrm855. PMID   12094216. S2CID   12384262.
  16. Bai J, Tucker WC, Chapman ER (January 2004). "PIP2 increases the speed of response of synaptotagmin and steers its membrane-penetration activity toward the plasma membrane". Nature Structural & Molecular Biology. 11 (1): 36–44. doi:10.1038/nsmb709. PMID   14718921. S2CID   1311311.
  17. Gaffaney, Jon D.; Dunning, F. Mark; Wang, Zhao; Hui, Enfu; Chapman, Edwin R. (2008-11-14). "Synaptotagmin C2B domain regulates Ca2+-triggered fusion in vitro: critical residues revealed by scanning alanine mutagenesis". The Journal of Biological Chemistry. 283 (46): 31763–31775. doi: 10.1074/jbc.M803355200 . ISSN   0021-9258. PMC   2581593 . PMID   18784080.
  18. 1 2 Kikuma, Koto; Li, Xiling; Kim, Daniel; Sutter, David; Dickman, Dion K. (November 2017). "Extended Synaptotagmin Localizes to Presynaptic ER and Promotes Neurotransmission and Synaptic Growth in Drosophila". Genetics. 207 (3): 993–1006. doi:10.1534/genetics.117.300261. ISSN   1943-2631. PMC   5676231 . PMID   28882990.
  19. Wu, Dick; Bacaj, Taulant; Morishita, Wade; Goswami, Debanjan; Arendt, Kristin L.; Xu, Wei; Chen, Lu; Malenka, Robert C.; Südhof, Thomas C. (20 April 2017). "Postsynaptic synaptotagmins mediate AMPA receptor exocytosis during LTP". Nature. 544 (7650): 316–321. Bibcode:2017Natur.544..316W. doi:10.1038/nature21720. ISSN   1476-4687. PMC   5734942 . PMID   28355182.
  20. Sreetama, S. C.; Takano, T.; Nedergaard, M.; Simon, S. M.; Jaiswal, J. K. (April 2016). "Injured astrocytes are repaired by Synaptotagmin XI-regulated lysosome exocytosis". Cell Death and Differentiation. 23 (4): 596–607. doi:10.1038/cdd.2015.124. ISSN   1476-5403. PMC   4986631 . PMID   26450452.
  21. Shen, Wei; Wang, Qiu-Wen; Liu, Yao-Nan; Marchetto, Maria C.; Linker, Sara; Lu, Si-Yao; Chen, Yun; Liu, Chuihong; Guo, Chongye; Xing, Zhikai; Shi, Wei (25 February 2020). "Synaptotagmin-7 is a key factor for bipolar-like behavioral abnormalities in mice". Proceedings of the National Academy of Sciences of the United States of America. 117 (8): 4392–4399. doi:10.1073/pnas.1918165117. ISSN   1091-6490. PMC   7049155 . PMID   32041882.

Related Research Articles

<span class="mw-page-title-main">Chemical synapse</span> Biological junctions through which neurons signals can be sent

Chemical synapses are biological junctions through which neurons' signals can be sent to each other and to non-neuronal cells such as those in muscles or glands. Chemical synapses allow neurons to form circuits within the central nervous system. They are crucial to the biological computations that underlie perception and thought. They allow the nervous system to connect to and control other systems of the body.

<span class="mw-page-title-main">Exocytosis</span> Active transport and bulk transport in which a cell transports molecules out of the cell

Exocytosis is a form of active transport and bulk transport in which a cell transports molecules out of the cell. As an active transport mechanism, exocytosis requires the use of energy to transport material. Exocytosis and its counterpart, endocytosis, are used by all cells because most chemical substances important to them are large polar molecules that cannot pass through the hydrophobic portion of the cell membrane by passive means. Exocytosis is the process by which a large amount of molecules are released; thus it is a form of bulk transport. Exocytosis occurs via secretory portals at the cell plasma membrane called porosomes. Porosomes are permanent cup-shaped lipoprotein structure at the cell plasma membrane, where secretory vesicles transiently dock and fuse to release intra-vesicular contents from the cell.

<span class="mw-page-title-main">Synaptic vesicle</span> Neurotransmitters that are released at the synapse

In a neuron, synaptic vesicles store various neurotransmitters that are released at the synapse. The release is regulated by a voltage-dependent calcium channel. Vesicles are essential for propagating nerve impulses between neurons and are constantly recreated by the cell. The area in the axon that holds groups of vesicles is an axon terminal or "terminal bouton". Up to 130 vesicles can be released per bouton over a ten-minute period of stimulation at 0.2 Hz. In the visual cortex of the human brain, synaptic vesicles have an average diameter of 39.5 nanometers (nm) with a standard deviation of 5.1 nm.

<span class="mw-page-title-main">SNARE protein</span> Protein family

SNARE proteins – "SNAPREceptors" – are a large protein family consisting of at least 24 members in yeasts, more than 60 members in mammalian cells, and some numbers in plants. The primary role of SNARE proteins is to mediate the fusion of vesicles with the target membrane; this notably mediates exocytosis, but can also mediate the fusion of vesicles with membrane-bound compartments. The best studied SNAREs are those that mediate the release of synaptic vesicles containing neurotransmitters in neurons. These neuronal SNAREs are the targets of the neurotoxins responsible for botulism and tetanus produced by certain bacteria.

<span class="mw-page-title-main">SNAP25</span> Protein-coding gene in the species Homo sapiens

Synaptosomal-Associated Protein, 25kDa (SNAP-25) is a Target Soluble NSF (N-ethylmaleimide-sensitive factor) Attachment Protein Receptor (t-SNARE) protein encoded by the SNAP25 gene found on chromosome 20p12.2 in humans. SNAP-25 is a component of the trans-SNARE complex, which accounts for membrane fusion specificity and directly executes fusion by forming a tight complex that brings the synaptic vesicle and plasma membranes together.

<span class="mw-page-title-main">C2 domain</span>

A C2 domain is a protein structural domain involved in targeting proteins to cell membranes. The typical version (PKC-C2) has a beta-sandwich composed of 8 β-strands that co-ordinates two or three calcium ions, which bind in a cavity formed by the first and final loops of the domain, on the membrane binding face. Many other C2 domain families don't have calcium binding activity.

<span class="mw-page-title-main">Complexin</span>

Complexin (also known as synaphin) refers to a one of a small set of eukaryotic cytoplasmic neuronal proteins which binds to the SNARE protein complex (SNAREpin) with a high affinity. These are called synaphin 1 and 2. In the presence of Ca2+, the transport vesicle protein synaptotagmin displaces complexin, allowing the SNARE protein complex to bind the transport vesicle to the presynaptic membrane.

<span class="mw-page-title-main">STX1A</span> Protein-coding gene in the species Homo sapiens

Syntaxin-1A is a protein that in humans is encoded by the STX1A gene.

<span class="mw-page-title-main">SYT1</span> Protein-coding gene in the species Homo sapiens

Synaptotagmin-1 is a protein that in humans is encoded by the SYT1 gene.

<span class="mw-page-title-main">RIMS1</span> Gene of the species Homo sapiens

Regulating synaptic membrane exocytosis protein 1 is a protein that in humans is encoded by the RIMS1 gene.

<span class="mw-page-title-main">DNAJC5</span> Protein-coding gene in the species Homo sapiens

DnaJ homolog subfamily C member 5, also known as cysteine string protein or CSP is a protein, that in humans encoded by the DNAJC5 gene. It was first described in 1990.

<span class="mw-page-title-main">DOC2B</span> Protein-coding gene in the species Homo sapiens

Double C2-like domain-containing protein beta is a protein that in humans is encoded by the DOC2B gene.

<span class="mw-page-title-main">Axon terminal</span> Nerve fiber part

Axon terminals are distal terminations of the branches of an axon. An axon, also called a nerve fiber, is a long, slender projection of a nerve cell that conducts electrical impulses called action potentials away from the neuron's cell body in order to transmit those impulses to other neurons, muscle cells or glands. In the central nervous system, most presynaptic terminals are actually formed along the axons, not at their ends.

The ribbon synapse is a type of neuronal synapse characterized by the presence of an electron-dense structure, the synaptic ribbon, that holds vesicles close to the active zone. It is characterized by a tight vesicle-calcium channel coupling that promotes rapid neurotransmitter release and sustained signal transmission. Ribbon synapses undergo a cycle of exocytosis and endocytosis in response to graded changes of membrane potential. It has been proposed that most ribbon synapses undergo a special type of exocytosis based on coordinated multivesicular release. This interpretation has recently been questioned at the inner hair cell ribbon synapse, where it has been instead proposed that exocytosis is described by uniquantal release shaped by a flickering vesicle fusion pore.

<span class="mw-page-title-main">Syntaxin</span> Group of proteins

Syntaxins are a family of membrane integrated Q-SNARE proteins participating in exocytosis.

Vesicle fusion is the merging of a vesicle with other vesicles or a part of a cell membrane. In the latter case, it is the end stage of secretion from secretory vesicles, where their contents are expelled from the cell through exocytosis. Vesicles can also fuse with other target cell compartments, such as a lysosome. Exocytosis occurs when secretory vesicles transiently dock and fuse at the base of cup-shaped structures at the cell plasma membrane called porosome, the universal secretory machinery in cells. Vesicle fusion may depend on SNARE proteins in the presence of increased intracellular calcium (Ca2+) concentration.

<span class="mw-page-title-main">Synapsin I</span> Protein-coding gene in the species Homo sapiens

Synapsin I, is the collective name for Synapsin Ia and Synapsin Ib, two nearly identical phosphoproteins that in humans are encoded by the SYN1 gene. In its phosphorylated form, Synapsin I may also be referred to as phosphosynaspin I. Synapsin I is the first of the proteins in the synapsin family of phosphoproteins in the synaptic vesicles present in the central and peripheral nervous systems. Synapsin Ia and Ib are close in length and almost the same in make up, however, Synapsin Ib stops short of the last segment of the C-terminal in the amino acid sequence found in Synapsin Ia.

<span class="mw-page-title-main">Active zone</span>

The active zone or synaptic active zone is a term first used by Couteaux and Pecot-Dechavassinein in 1970 to define the site of neurotransmitter release. Two neurons make near contact through structures called synapses allowing them to communicate with each other. As shown in the adjacent diagram, a synapse consists of the presynaptic bouton of one neuron which stores vesicles containing neurotransmitter, and a second, postsynaptic neuron which bears receptors for the neurotransmitter, together with a gap between the two called the synaptic cleft. When an action potential reaches the presynaptic bouton, the contents of the vesicles are released into the synaptic cleft and the released neurotransmitter travels across the cleft to the postsynaptic neuron and activates the receptors on the postsynaptic membrane.

<span class="mw-page-title-main">Synaptic stabilization</span> Modifying synaptic strength via cell adhesion molecules

Synaptic stabilization is crucial in the developing and adult nervous systems and is considered a result of the late phase of long-term potentiation (LTP). The mechanism involves strengthening and maintaining active synapses through increased expression of cytoskeletal and extracellular matrix elements and postsynaptic scaffold proteins, while pruning less active ones. For example, cell adhesion molecules (CAMs) play a large role in synaptic maintenance and stabilization. Gerald Edelman discovered CAMs and studied their function during development, which showed CAMs are required for cell migration and the formation of the entire nervous system. In the adult nervous system, CAMs play an integral role in synaptic plasticity relating to learning and memory.

<span class="mw-page-title-main">Ferlins</span> Protein family

Ferlins are an ancient protein family involved in vesicle fusion and membrane trafficking. Ferlins are distinguished by their multiple tandem C2 domains, and sometimes a FerA and a DysF domain. Mutations in ferlins can cause human diseases such as muscular dystrophy and deafness. Abnormalities in expression of myoferlin, a human ferlin protein, is also directly associated with higher mortality rate and tumor recurrence in several types of cancer, including pancreatic, colorectal, breast, cervical, stomach, ovarian, cervical, thyroid, endometrial, and oropharyngeal squamous cell carcinoma. In other animals, ferlin mutations can cause infertility.