MOLECULAR & CELLULAR NEUROBIOLOGY 
Master Course Cognitive Neuroscience - Radboud University, Nijmegen

 

INDEX

INTRODUCTION CELLS AND WITHIN CELLS IN A NUTSHELL GENOMICS MOLECULAR BIOLOGICAL RESEARCH METHODOLOGY NEURODEVELOPMENT  

 

Chapter 2:  Cells and within cells

 

Cells

DNA and genes

Translation

Receptor Mechanisms

 

    Neurons

   More on DNA

   Proteins, Protein Structure and Protein Analysis

   Ion channel receptors

 

    Glia

   Epigenetics

   Protein folding in the cell

   Tyrosine kinase receptors

Within cells

   Transcription

   Post-translational modifications of proteins

   G-protein-coupled receptors
   Amino ac, Carbohydr, Lipids and Nucleic ac

   Noncoding RNAs

   Protein degradation in the cell - Autophagy

   G-proteins

   Membranes and Membrane Proteins

   miRNAs and the brain

   Protein secretion / Secretory pathway

   Transcription and signalling

   The Exctracellular Matrix

       Transcription factor receptors

 

 

Protein secretion / Secretory pathway

The section 'Protein secretion/Secretory pathway' deals with a number of cellular subcompartments (notably the endoplasmic reticulum and Golgi network) and therefore represents a long section with many aspects. For those interested, more specific details are given in blocks of texts in italic (these parts in italic do not belong to the exam material).

 
The endoplasmic reticulum (ER) 

The first compartment of the secretory pathway is the endoplasmic reticulum (ER). The ER is a continuous membrane system, but consists of various domains that perform different functions. Structurally distinct domains of this organelle include the nuclear envelope (NE), the rough and smooth ER (RER and SER, respectively), and the regions that contact other organelles. The establishment of these domains and the targeting of proteins to them are understood to varying degrees. Despite its complexity, the ER is a dynamic structure. The ER has many different functions. These include the translocation of proteins (such as secretory proteins) across the ER membrane; the integration of proteins into the membrane; the folding and modification of proteins in the ER lumen; the synthesis of phospholipids and steroids on the cytosolic side of the ER membrane; and the storage of calcium ions in the ER lumen (used for intracellular signaling in response to mitogenic and growth factor signal transduction pathways) and their regulated release into the cytosol.

The morphological differences between RER and SER allow these two regions of the ER to be distinguished visually; for example, the SER is often more convoluted than RER, and the RER tends to be more granular in texture. These differences in appearance may be directly related to the presence of bound ribosomes on the RER as there is some evidence that this affects ER structure. Ultimately, however, the distinction between the two must be explained by differences in membrane protein composition. Most membrane proteins are shared between RER and SER (general ER proteins), but several proteins involved in translocation or processing of newly synthesized proteins are enriched in RER, as shown by the fractionation of liver cells. Since protein translocation is essential for all eukaryotic cells, they all have RER. One type of SER that is also found in all cells is the transitional ER. It is involved in packaging proteins for transport from the ER to the Golgi and is enriched in proteins required for this process. However, SER is abundant only in certain cell types, such as in steroid-synthesizing cells, liver cells, neurons and muscle cells. The primary activities of the SER are very different in each of these cell types. The SER acts as an overflow site to house upregulated enzymes, and as these enzymes vary, it is also a cell type-specific suborganelle. The relative abundance of RER and SER found among different cell types correlates with their functions. For example, cells that secrete a large percentage of their synthesized proteins contain mostly RER.The ER is closely associated with essentially all other organelles in the cell. These include the plasma membrane, Golgi, vacuoles, mitochondria, peroxisomes, late endosomes and lysosomes. In order to travel along the secretory pathway and eventually reach their appropriate cellular destinations, newly synthesized secreted and membrane-bound proteins must fold and assemble correctly (the ER is a major protein folding compartment in a eukaryotic cell). Failure to do so results in their retention in the ER and eventual degradation. The proper conformational maturation of nascent secretory pathway proteins in the ER is both aided and monitored by a number of so-called chaperones and folding enzymes in a complex process termed ER quality control. A multitude of post-translational modifications occur in the ER, including N-linked glycosylation, disulfide bond formation, lipidation, hydroxylation and oligomerization.

 

         Electron microscope view: rough ER (RER) with its attached ribosomes; Mit: mitochondrion.

.

Electron microscope view: smooth ER (SER) is shorter and thicker than RER, and associated with lipid synthesis; Nu: nucleus; NuPr: nuclear pore.

Protein entry and maturation in the ER  

Transport systems which act in a co- or posttranslational mode ensure that all proteins are targeted to the correct location within the cell. In eukaryotes, the signal recognition particle (SRP) mediates the transport of secretory and membrane proteins to the ER (Figures 1A and B). Signal sequences of target proteins are specifically recognized by SRP as they emerge from the ribosome. Typical signal sequences have a 9–12-residue-long hydrophobic stretch in the middle. Although for binding to the SRP the hydrophobic part of the signal peptide is crucial, the flanking regions may also contribute to the interaction. Thus, the SRP and its membrane-bound receptor (SR) deliver membrane proteins and secretory proteins to the translocation channel in the ER. The general outline of the SRP pathway is conserved in all three kingdoms of life.

About 20% of all proteins encoded by the human genome are predicted to be secretory proteins. Targeted by the signal sequence, polypeptides enter the secretory pathway in an unfolded state. In the ER, a specialized environment controls posttranslational modifications and allows newly imported polypeptides to assume their native structures and to assemble into multimeric complexes. Because solvent-exposed, hydrophobic segments present in unfolded or partially folded proteins tend to aggregate, BiP (a member of the Hsp70 family) and other ER-resident chaperones bind to such hydrophobic patches and preserve the folding competence of the nascent chain.

             

   

 

 

 

 

 

 

Figure 1A. Protein synthesis in the ER.

 

 

 

 

 

  Figure 1B. Schematic representation of the SRP cycle with four general steps. Protein transport starts with the recognition of the signal peptide (‘SP’) on the ribosome. The SRP/ribosome-nascent-chain complex is then binding to the SRP receptor (SR). For the formation of a stable SRP/SR complex, GTP has to be present in both SRP and the SR. The signal peptide is transferred from SRP to the translocation channel. GTP hydrolysis in both, SRP and SR, leads to the dissociation of the SRP/SR complex. SRP is represented by a red sphere, the SRP receptor is in pink, the translocation channel in green and the ribosome in yellow. The signal peptide at the N-terminus of a nascent polypeptide is shown in blue.

 

 

Several enzymatic activities alter the secondary structure of the polypeptide chain during and after import. Peptidylprolyl-isomerase (PPIs/immunophilins) catalyzes the cis–trans isomerization of proline residues, and  oxidoreductases, such as protein disulfide isomerase (PDI) and Erp57, control disulfide bond formation between correct pairs of cysteine residues.

A multitude of post-translational modifications occur in the ER: N-linked glycosylation, disulfide bond formation, lipidation, hydroxylation, oligomerization, etc. N-linked glycosylation is one of the most common posttranslational protein modifications and is .initiated by transfer of a core oligosaccharide to consensus Asn-X-Ser/Thr residues in the polypeptide chain. The covalent attachment of hydrophilic oligosaccharides serves several purposes in protein folding, assembly and trafficking: first, due to the hydrophilic nature of carbohydrates, glycosylation increases the solubility of glycoproteins and defines the attachment area for the surface of the protein. Second, due to their large hydrated volume oligosaccharides shield the attachment area from surrounding proteins. Third, oligosaccharides interact with the peptide backbone and stabilize its conformation. Lastly, sequential trimming of sugar residues is monitored by a lectin machinery to report on the folding status of the protein.

The calnexin/calreticulin cycle is one arm of a complex process termed the quality-control machinery in the ER that monitors protein conformations and dictates whether a molecule is exported to the Golgi or targeted for ER-associated degradation (ERAD). Thus, proteins that have ultimately failed ER quality control are degraded to prevent their accumulation in the ER, which might either titrate out the components of the chaperone systems or form large insoluble aggregates that would be toxic to the cell. ERAD is conserved from lower eukaryotes like yeast to mammals. Both malfolded proteins and excess subunits of multimeric proteins are retrotranslocated or dislocated back into the cytosol, which appears to be similar to the translocon used to enter the ER lumen. This retro-translocation process is usually coupled with ubiquitination, which occurs at the cytosolic surface of the ER membrane. Ubiquitin (Ub) is a highly conserved small protein that is universally expressed in eukaryotic cells. Ubiquitination of substrates is a multi-step process. Thus, if improperly folded, the protein is  retrogradely translocated to the cytosol for degradation by the proteasome (see Figure 2 and also under “Protein degradation in the cell). Puzzling is the fact that several parallel pathways appear to exist that are capable of extracting proteins from the ER and delivering them to the proteasome for destruction. Given the complexity of the system and the medical importance of this process, the detailed study of ERAD will remain an interesting topic for the next years.

Thus, in order to travel along the secretory pathway and eventually reach their appropriate cellular destinations, newly synthesized secreted and membrane-bound proteins must fold and assemble correctly. Failure to do so results in their retention in the ER and eventual degradation. The proper conformational maturation of nascent secretory pathway proteins is both aided and monitored by a number of ER chaperones and folding enzymes in the ER quality control process mentioned above. The transport of newly synthesized polypeptides that have not yet folded completely, as well as those proteins that have folded incorrectly, are retained in the ER via their interactions with the chaperones (Figure 2; see also under “Protein folding in the cell”).

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Figure 2. Schematic illustration of ER-resident chaperone functions under non-stress conditions including (1) facilitating co-translational translocation; (2) helping protein folding; (3) facilitating retro-translocation and ERAD; and (4) contributing to ER luminal calcium storage.

 
 
Storage of cellular calcium in the ER    

A major function of the ER is to store calcium that is used for intracellular signaling in response to mitogenic and growth factor signal transduction pathways. Calcium is pumped into the ER via the action of ER-localized, transmembrane ATPases known as SERCA pumps in mammalian cells. In the ER, calcium is bound through both high affinity and low affinity interactions to a number of resident ER proteins including calnexin, calreticulin, GRP94, BiP, and CaBP1. Sustained ER calcium levels are essential for normal protein folding in this organelle, as drugs like thapsigargin, which interferes with the action of the ER calcium ATPase, and A23187, which depletes ER calcium, result in dramatic and rapid activation of the ER stress or unfolded protein response (UPR).  

ER stress   

In its broadest definition, stress is the response of any system to perturbations of its normal state. For a cell or organism these can be either life-enhancing changes, e.g. feeding, or life-threatening changes, e.g. starvation. To apply this definition of stress to an organelle, e.g. the ER, we have to address the following questions: what are the physiological functions of the ER and how are they perturbated? Furthermore, we have to understand how these perturbations are sensed and how signals are transduced to initiate countermeasures to restore the original state. Only parts of these questions have been answered above and will be dealt with here.

In eukaryotic cells, the ER is the first compartment in the secretory pathway. As discussed above, it is responsible for the synthesis, modification and delivery of proteins to their proper target sites within the secretory pathway and the extracellular space. All secretory proteins enter the secretory pathway through the ER. In addition, the ER is the site for the synthesis of sterols and lipids. In lower eukaryotes, a major portion of the cell wall is synthesized in the ER. Disruption of any of these processes causes ER stress. Historically, the focus is on ER stress caused by disruption of protein folding, and little is currently known about ER stress caused, for example, by aberrations in lipid metabolism, or disruption of cell wall biogenesis. Proof of principle experiments established that expression of mutant, folding-incompetent proteins causes ER stress and an ER stress response (unfolded protein response, UPR). This is the biochemical basis for many ER storage diseases, in which folding-incompetent proteins accumulate in the ER. In vivo protein folding requires a complex ER-resident protein folding machinery. Exhaustion of the capacity of this protein folding machinery by over-expression of wild-type proteins, e.g. blood coagulation factor VIII, or antithrombin III results in the accumulation of unfolded, aggregated proteins in the ER and activation of the UPR. Recently, many physiological conditions were identified in which the demand on the ER-resident protein folding machinery exceeds its capacity, e.g. differentiation of B-cells into plasma cells, a cell type highly  specialized in secretion. 

Two simple adaptive mechanisms are employed to bring the folding capacity of the ER and its unfolded protein burden into line and return the ER to its normal physiological state: (1) upregulation of the folding capacity of the ER through induction of ER-resident molecular chaperones and foldases, and an increase in the size of the ER; the transcriptional up-regulation of ER chaperones is the hallmark of the ER stress response and occurs in all eukaryotic organisms, and (2) down-regulation of the biosynthetic load of the ER through shut-off of protein synthesis on a transcriptional and translational level, and increased clearance of unfolded proteins from the ER through upregulation of ERAD.

Conformational diseases are caused by mutations altering the folding pathway or final conformation of a protein. Many conformational diseases are caused by mutations in secretory proteins and reach from metabolic diseases, e.g. diabetes, to developmental and neurological diseases, e.g. Alzheimer’s disease (see also below).  

In terms of functions of the ER chaperones during the stress response, it would appear in many cases that they do the same thing as during normal physiological conditions but perhaps more so. A major function of the ER chaperones is to promote protein folding by preventing misfolding or aggregation. During conditions of ER stress, alterations in the ER environment can profoundly affect the folding of many proteins. This includes proteins that bind to the BiP chaperone system, as well as those that bind only to the calnexin/calreticulin system. Although BiP appears to be the sole chaperone that is monitored by the cell to sense ER stress, many of the chaperones are coordinately up-regulated. Thus, the main function of the increased levels of ER chaperones is to bind to unfolded proteins, prevent them from aggregating, and to aid and monitor their refolding if normal physiological conditions are restored to the ER. If the stress is not resolved rapidly, many unfolded proteins will be targeted for ERAD as one way to decrease the load of malfolded proteins that accumulate in the ER. So conceptually, UPR and ERAD are partially overlapping means to the same end, and it makes sense for the cell to regulate ERAD during UPR. Finally, if the ER stress conditions cannot be alleviated, in order to ultimately protect the organism, presumably by eliminating unhealthy or infected cellsan apoptotic pathway is activated, which involves the ER-localized caspase-12 protein.

Thus, a vast number of studies have revealed a role for ER chaperones in contributing to or controlling all of the major functions of the ER, including translocation of nascent polypeptide chains, folding and assembly of secretory pathway proteins, monitoring the success of this operation, and identifying those proteins that fail this quality control and targeting them for degradation. Clearly these functions of ER chaperones would be even more critical during conditions of ER stress. In addition, they play a role in storing calcium in the ER, which is important in cellular signal transduction pathways, and which also may play a role in initiating apoptosis when ER stress is sustained and induces loss of calcium from the ER. Finally, the ER chaperone BiP is central to the cellular mechanisms to sense alterations in the ER. 

 

Genetic and biochemical dissection of the secretory pathway   

In the late 1970s the molecular mechanisms that underlie vesicular transport were elucidated by reducing vesicular transport to a set of elementary biochemical reactions: Schekman and colleagues isolated and studied temperature-sensitive "sec" mutants of the yeast Saccharomyces cerevisiae that were defective in protein secretion, while Rothman and colleagues devised an ingenious cell-free assay to measure protein transport between cisternae of the mammalian Golgi complex (in vitro reconstitution). From these studies, it was concluded that yeast and mammals share a conserved vesicular transport machinery, which can be dissected using both genetic and biochemical tools. The results have produced a detailed molecular picture of the mechanisms of trafficking in the secretory pathway and the related endocytic and vacuolar/lysosomal targeting pathways. Central to these mechanisms are the two most critical events in the lifetime of a transport vesicle, namely budding and fusion.

ER-to-Golgi transport     

Each of the transport steps connecting ER and Golgi, and the Golgi compartments is unidirectional, and energy- and GTP dependent. Small GTP-binding proteins (20-30 kDa, e.g. the rab proteins) are distributed throughout the secretory pathway with distinctive subcellular locations and regulate vesicular trafficking between the subcompartments. In addition, the traffic is regulated by integral membrane proteins (vesicular soluble NSF attachment receptors, v-SNAREs, and target t-SNAREs; NSF=N-Ethylmaleimide-sensitive factor, v=vesicular, t=target) for docking the vesicle to the membrane of the acceptor compartment and soluble proteins (NSF and SNAPs) to initiate vesicle fusion. In this vesicular model of protein transport, all anterograde (forward) and retrograde (backward) intra-Golgi steps involve transport only via vesicles. Alternatively, a nonvesicular transport mechanism may be present in which Golgi cisternae form at the cis-face of the stack, probably by VTC fusion, and then progressively mature into trans-cisternae (cisternal maturation model). In this model, cisternae (or intermittent tubular continuities) carry secretory cargo through the stack in the anterograde direction, while vesicles transport Golgi enzymes in the retrograde direction, allowing cisternal maturation to occur by progressive uptake of material from older stacks. At the TGN, the cisternae ultimately disintegrate and evolve into a collection of secretory vesicles, including immature secretory granules.  

Vesicular transport from the ER to the Golgi complex constitutes the initial step in protein secretion. So-called COPII-coated vesicles (see below) mediate the export of newly synthesized proteins from the ER, and this transport step is coupled with COPI-mediated retrograde traffic to form a transport circuit that supports the compositional asymmetry of the ER-Golgi system. Biochemical and structural studies have advanced our understanding of the mechanisms that control vesicle formation and cargo-protein capture. Recent work has highlighted the function of transitional ER regions in specifying the location of COPII budding. COPI coat components traffic primarily from the Golgi to the ER and between Golgi cisternae. COPII-coated vesicles traffic from the ER to the Golgi. COP I and COP II coat proteins thus direct protein and membrane trafficking in between early compartments of the secretory pathway in eukaryotic cells. These coat proteins perform the dual, essential tasks of selecting appropriate cargo proteins and deforming the lipid bilayer of appropriate donor membranes into buds and vesicles. COP II proteins are required for selective export of newly synthesized proteins from the ER. COP I proteins mediate a retrograde transport pathway that selectively recycles proteins from the cis-Golgi complex to the ER. Additionally, COP I coat proteins have complex functions in intra-Golgi trafficking and in maintaining the normal structure of the mammalian interphase Golgi complex. 

Selective cargo export from the ER is brought about by the budding of COPII vesicles. While the main structural components of the COPII coat have been identified and characterized, the regulatory event(s) promoting COPII vesicle biogenesis and cargo selection still remains largely unknown. COPII vesicle biogenesis from the ER is critical for the transit of all cargo molecules to the Golgi. This process depends on vesicle budding from the ER membrane through a cycle of coat polymerisation. Following scission, unknown events trigger coat disassembly and targeting of the vesicle to the Golgi.

Proteins destined for the Golgi exit the ER at specialized ER exit sites (ERES) defined by the presence of COPII coats. The close apposition of ERES to the cis-face of the Golgi stack led to the "cisternal maturation" model for Golgi biogenesis, in which new cis-cisterna form from membranes derived from the adjacent ERES. The cisterna matures as it moves in the trans-direction through the stack, by recycling components in the cis-direction. An alternative "vesicular" model proposes that each Golgi cisterna is a stable compartment, and receives cargo by vesicular traffic. The cisternal maturation model predicts that Golgi are always spatially associated with ERES. The vesicular model allows Golgi localization to be independent of ERES.

 

The coated vesicle budding hypothesis  

The vesicle-transport hypothesis for protein trafficking was based on early electron microscopy studies. Vesicular transport intermediates bud from a donor organelle and then fuse with an acceptor organelle. Budding intermediates were initially identified by their electron-dense ‘coats’ and were found on the plasma membrane and intracellular organelles. Three major classes of these coated vesicles have now been purified: COPI- (coatomer) and COPII-coated vesicles (where COP stands for coat protein complex) and clathrin-coated vesicles. Coat components are needed for generation of highly curved membrane areas, recruitment of cargo (and exclusion of non-cargo proteins/lipids), vesicle scission and uncoating factor recruitment.

Just as a SNARE-based fusion underpins the hypothesis of a common mechanism for all membrane fusion, so a coat-based budding hypothesis has emerged as a common theme in vesicle budding. Clathrin-coated vesicles are named after the protein that self-polymerises into a lattice around these vesicles as they bud from the plasma membrane, trans-Golgi network (TGN) and endosomes. For all clathrin-coated vesicles, clathrin is the central organiser. It concentrates cargo adaptors, leading to a diverse protein and lipid load in the forming vesicle, and its polymerisation into a curved lattice stabilises the nascent membrane bud as it forms. The main cargo adaptor proteins characterised to date are the classical adaptor protein (AP) complexes, AP1, AP2, AP3 and AP4. Most AP complexes adapt and/or link clathrin to selected membrane cargo and lipids, and they also bind accessory proteins that regulate coat assembly and disassembly.

A long-recognised commonality between COPI, COPII and many clathrin budding pathways is that small G-proteins are required for membrane recruitment. COPI components and the clathrin adaptors (AP1, AP3, AP4 and GGAs) are recruited by Arfs (ADP ribosylation factor substrates), whereas COPII depends on the Arf-related protein Sar1. COPI and COPII coats all have the equivalent of a clathrin and an adaptor subcomplex.

Vesicle budding and fusion 

Genetic and biochemical analyses of the secretory pathway have produced a detailed picture of the molecular mechanisms involved in selective cargo transport between organelles. This transport occurs by means of vesicular intermediates that bud from a donor compartment and fuse with an acceptor compartment. Vesicle budding and cargo selection are mediated by protein coats, while vesicle targeting and fusion depend on a machinery that includes the SNARE proteins. Precise regulation of these two aspects of vesicular transport ensures efficient cargo transfer while preserving organelle identity.

The vesicular transport hypothesis

The stage was set over 30 years ago by the work of George Palade and colleagues on protein secretion. This work established that newly synthesized secretory proteins pass through a series of membrane-enclosed organelles, including the ER, the Golgi complex and secretory granules, on their way to the extracellular space. Proteins destined for residence at the plasma membrane, endosomes, or lysosomes share the early stations of this pathway (i.e., the ER and the Golgi complex) with secretory proteins. Importantly, the secretory proteins are often found within small, membrane-enclosed vesicles interspersed among the major organelles of the pathway. Such observations inspired the vesicular transport hypothesis, which states that the transfer of cargo molecules between organelles of the secretory pathway is mediated by shuttling transport vesicles. According to this hypothesis, vesicles bud from a "donor" compartment ("vesicle budding") by a process that allows selective incorporation of cargo into the forming vesicles while retaining resident proteins in the donor compartment ("protein sorting"). The vesicles are subsequently targeted to a specific "acceptor" compartment ("vesicle targeting"), into which they unload their cargo upon fusion of their limiting membranes ("vesicle fusion"). An updated representation of the steps of vesicular transport is shown in Figure 3. The processes of budding and fusion are iterated at the consecutive transport steps until the cargo reaches its final destination within or outside the cell. To balance this forward movement of cargo, organelle homeostasis requires the retrieval of transport machinery components and escaped resident proteins from the acceptor compartments back to the corresponding donor compartments ("retrograde transport"), a process that is also proposed to occur by vesicular transport. All of these steps are tightly regulated and balanced so that a large amount of cargo can flow through the secretory pathway without compromising the integrity and steady-state composition of the constituent organelles.

 

Figure 3. Steps of vesicle budding and fusion. (1) Initiation of coat assembly. The membrane-proximal coat components (blue) are recruited to the donor compartment by binding to a membrane-associated GTPase (red) and/or to a specific phosphoinositide. Transmembrane cargo proteins and SNAREs begin to gather at the assembling coat. (2) Budding. The membrane-distal coat components (green) are added and polymerize into a mesh-like structure. Cargo becomes concentrated and membrane curvature increases. (3) Scission. The neck between the vesicle and the donor compartment is severed either by direct action of the coat or by accessory proteins. (4) Uncoating. The vesicle loses its coat due to various events including inactivation of the small GTPase, phosphoinositide hydrolysis, and the action of uncoating enzymes. Cytosolic coat proteins are then recycled for additional rounds of vesicle budding. (5) Tethering. The “naked” vesicle moves to the acceptor compartment, possibly guided by the cytoskeleton, and becomes tethered to the acceptor compartment by the combination of a GTP bound Rab and a tethering factor. (6) Docking. The v- and t-SNAREs assemble into a four-helix bundle. (7) This “trans-SNARE complex” promotes fusion of the vesicle and acceptor lipid bilayers. Cargo is transferred to the acceptor compartment, and the SNAREs are recycled.

 

 

 

 

 

 

 

Role of protein coats in vesicle budding and cargo selection

The budding of transport vesicles and the selective incorporation of cargo into the forming vesicles are both mediated by protein coats. These coats are supramolecular assemblies of proteins that are recruited from the cytosol to the nascent vesicles. The coats deform flat membrane patches into round buds, eventually leading to the release of coated transport vesicles. The coats also participate in cargo selection by recognizing sorting signals present in the cytosolic domains of transmembrane cargo proteins. Vesicle budding and cargo selection at different stages of the exocytic and endocytic pathways are mediated by different coats and sorting signals. The first coats to be identified and characterized contained a scaffold protein, clathrin, as their main constituent. Clathrin coats were initially assumed to participate in most, if not all, vesicular transport steps within the cell. However, later studies demonstrated that the function of these coats was restricted to post-Golgi locations including the plasma membrane, the trans-Golgi network (TGN), and endosomes. A major discovery was the existence of non-clathrin coats that mediate vesicular transport in the early secretory pathway. One of these coats, COPII, is now known to mediate export from the ER to either the ER-Golgi intermediate compartment (ERGIC) or the Golgi complex, while another coat, COPI, is involved in intra-Golgi transport and retrograde transport from the Golgi to the ER. Of the various protein coats that have been identified to date, COPII is one of the best understood. Apart from Sar1p, the subunits of the COPII coat are structurally distinct from those of the COPI and clathrin coats. The relative simplicity of COPII, as well as its unique role in ER export, have facilitated the analysis of its assembly and function.

Cargo selection

The majority of cargo proteins are actively concentrated in COPII-coated buds and vesicles prior to export from the ER. Most transmembrane cargo proteins exit the ER by binding directly to COPII, but some transmembrane and most soluble cargo proteins bind indirectly to COPII through transmembrane export receptors. Export receptors leave the ER together with their ligands, unload their cargo into the acceptor compartment, and recycle back to the ER. The sorting signals recognized by the COPII coat are found in the cytosolic domains of transmembrane cargo proteins. These signals are quite diverse. The involvement of so many different signals in the same sorting step implies the existence of either multiple binding sites on the same recognition protein or a family of recognition proteins. Both of these solutions have evolved for COPII. The diversity of signals and recognition modes explains the ability of COPII to package a wide variety of exported proteins.

Clathrin coats are considerably more complex than COPII and COPI. Clathrin and clathrin-adaptor complexes can polymerize into spherical, cage-like structures, as can COPII, indicating that these proteins have an intrinsic ability to sculpt buds and vesicles from membranes. Thus, the clathrin-adaptor complexes appear to perform the same basic functions as the COPII coats: cytosolic signal recognition and membrane deformation. However, the clathrin vesicle cycle involves additional classes of proteins that do not seem to operate during COPII vesicle formation.

Role of SNARE proteins in vesicle fusion

 

After a vesicle sheds its coat, it must be targeted to the appropriate acceptor compartment. The final step in a vesicle's existence is fusion with the acceptor membrane. Remarkably, the targeting and fusion reactions both rely on the same class of proteins.

The so-called SNARE hypothesis proposes that each type of transport vesicle carries a specific "v-SNARE" that binds to a cognate "t-SNARE" on the target membrane (Figure 4). This idea fits with the observations that various SNAREs localize to different intracellular compartments. SNAREs seem to perform two major functions. One function is to promote fusion itself: SNAREs form the conserved, essential core of the fusion machinery. It is likely that in all of the transport steps in the secretory and endocytic pathways, SNAREs perform the function of overcoming the energy barrier to fusion. The second major function of SNAREs is to help ensure the specificity of membrane fusion. Different v-/t-SNARE complexes form at different steps of intracellular transport. SNAREs cannot, however, be the only specificity determinants for membrane fusion because a given v-SNARE recycles and is therefore present in both anterograde and retrograde vesicles (Figure 5B). Like many biological processes, membrane fusion employs sequential, partially redundant mechanisms to achieve high fidelity.

Not surprisingly, a plethora of accessory components and regulatory reactions modulate the action of SNAREs. This modulation is important to prevent inappropriate events of SNARE complex formation. For instance, the putative Ca2+ sensor synaptotagmin interacts with SNAREs and promotes synaptic vesicle fusion in response to Ca2+ influx. Little is known about how SNAREs are targeted to specific organelles. For the few SNAREs that have been examined, targeting determinants are present in the transmembrane sequence, the cytosolic domain, or both. An important mechanism for SNARE localization is interaction with vesicle coats. For example, SNAREs involved in ER-to-Golgi transport must be packaged into COPII vesicles during ER export and then into COPI vesicles during retrieval from the Golgi.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Figure 4. Structure and function of SNAREs (A) Crystal structure of a synaptic trans- SNARE complex. (B) The SNARE cycle. A trans-SNARE complex assembles when a monomeric v-SNARE on the vesicle binds to n oligomeric t-SNARE on the target membrane, forming a stable four-helix bundle that promotes fusion. The result is a cis-SNARE complex in the fused membrane. a-SNAP binds to this complex and recruits NSF, which hydrolyzes ATP to dissociate the complex. Unpaired v-SNAREs can then be packaged again into vesicles.

 

 

The Golgi complex    

Once secretory proteins are properly folded and modified in the first compartment of the secretory pathway, the ER, the proteins move to specialized regions, the ER export sites, and are packaged in coated, irregularly shaped transport vesicles. Following uncoating, the transport vesicles form vesicular-tubular clusters (VTCs, also called the ER-Golgi intermediate compartment ERGIC) that move to the Golgi.

The scheme below depicts the compartments of the secretory pathway and two compartments associated with the secretory pathway, namely the lysosomal/vacuolar and endocytic pathways. Transport steps are indicated by arrows. Colors indicate the known or presumed locations of the transport vesicle coats COPII (blue), COPI (red), and clathrin (yellow/orange). Clathrin coats are heterogeneous and contain different adaptor and accessory proteins at different membranes. Only the function of COPII in ER export and of plasma membrane-associated clathrin in endocytosis are known with certainty. Less well understood are the exact functions of COPI at the ERGIC and Golgi complex and of clathrin at the TGN, early endosomes, and immature secretory granules. The pathway of transport through the Golgi stack is still being investigated, but is generally believed to involve a combination of COPI-mediated vesicular transport and cisternal maturation (see below). Additional coats or coat-like complexes exist but are not represented in this figure.

 

 

 

 

 

 

 

The Golgi apparatus modifies and sorts proteins for transport throughout the cell.The Golgi apparatus is often found in close proximity to the ER in cells. Protein cargo moves from the ER to the Golgi, is modified within the Golgi, and is then sent to various destinations in the cell, including the lysosomes and the cell surface.

 

 

The buds emerging from the ER thus become vesicles and are coated with COPII protein coats. After the vesicles lose their COPII coat, they merge with the VTCs (ERGIC) carrying the soluble and membrane proteins to the Golgi complex. Note that the vesicles are moving to contribute to the cis-Golgi network of vesicles and cisternae. The movement of these special transport vesicles is an energy-requiring process. If one blocks production of ATP, the transport will not happen. The drawing above shows how the ER forms vesicles (without ribosomes attached) that carry the newly synthesized proteins to the Golgi complex. The inside of the vesicle becomes continuous with the inside of the Golgi cisternae, so that protein groups pointing towards the inside, could eventually be directed to face the outside of the cell. Carbohydrate groups are attached and any subunits may be joined in these cisternae. The protein is then passed to the final region of the Golgi called the "trans face". There it is placed in vacuoles that bud from this region of the Golgi complex. These may be a certain size or density, characteristic of the cell itself. The vacuoles continue to condense the proteins and the final mature secretory granule is then moved to the membrane for secretion.

 

Transport of material in and out of the Golgi complex thus involves budding and fusion of vesicles. The cartoon shows that the membranes of each join and align themselves during the process so that the inside face remains in the lumen and the outside face remains towards the cytoplasm.

 

This electron micrograph illustrates the Golgi complex. The Golgi is curved with its trans face pointing away from the nucleus (top left) toward the cell periphery. The numerous vesicles in the area are transporting the proteins to and from cisternae.

 

 

Can proteins be transported from the Golgi back to the ER?  

Sometimes vital proteins needed in the ER are transported along with the other proteins to the Golgi complex. The Golgi complex has a mechanism for trapping them and sending them back to the ER. This cartoon shows the process. The protein destined for secretion is red (travels via the forward or anterograde pathway). The blue protein must remain in the ER. The ER has inserted a receptor protein on the membrane it sends to the Golgi complex in the transitional vesicles (the so-called KDEL receptor, shown in green).  These are retrograde vesicles and are therefore coated with COPI.  The KDEL receptor captures all of the  protein that carries the ER residency signal (the KDEL signal, dot in figure on the right, e.g., in BiP). Vesicles then bud from the Golgi complex and move back to the ER (retrograde pathway). The receptor can circulate and continue to return the proteins needed by the ER.

 

 

Golgi formation and cisternal maturation

There is much interest in understanding how the different Golgi cisternae are organized and differentiated. Two main models exist. In the “vesicular model” of protein transport (see figure above), all anterograde (forward) and retrograde (backward) intra-Golgi steps involve transport only via vesicles. Alternatively, a nonvesicular transport mechanism may be present in which Golgi cisternae form at the cis-face of the stack, probably by VTC fusion, and then progressively mature into trans-cisternae (“cisternal maturation model”; see figures on the right and below). In this model, cisternae (or intermittent tubular continuities) carry secretory cargo through the stack in the anterograde direction, while vesicles transport Golgi enzymes in the retrograde direction, allowing cisternal maturation to occur by progressive uptake of material from older stacks. At the TGN, the cisternae ultimately disintegrate and evolve into a collection of secretory vesicles, including immature secretory granules.                

  

    

                                                  

The Golgi complex controls trafficking of different types of proteins. Some are destined for secretion. Others are destined for the extracellular matrix. Finally, other proteins, such as lysosomal enzymes, may need to be sorted and sequestered from the remaining constituents because of their potential destructive effects. The figures below show the two types of secretory pathways. The regulated secretory pathway, as its name implies, is a pathway for proteins that requires a stimulus or trigger to elicit secretion. Some stimuli regulate synthesis of the protein as well as its release. The constitutive pathway allows for secretion of proteins that are needed outside the cell, like in the extracellular matrix. It does not require stimuli, although growth factors may enhance the process. Finally, the cartoons also show the packaging of lysosomes.

    

   

   

Golgi proteins    

The Golgi apparatus contains at least a thousand different types of integral and peripheral membrane proteins, perhaps more than any other intracellular organelle. The Golgi apparatus performs three major functions essential for growth, homeostasis and division of eukaryotic cells. First, it operates as a carbohydrate factory for the processing and modification of proteins and lipids moving through the secretory pathway. Second, it serves as a station for protein sorting and transport, receiving membrane from the ER and delivering it to the plasma membrane or other intracellular sites. Finally, it acts as a membrane scaffold onto which diverse signaling, sorting and cytoskeleton proteins adhere.

These distinct Golgi functions operate within a structure that is unique among subcellular organelles in many ways, including its composition as a stacked array of cisternae and connecting tubules/vesicles, its enormous diversity of protein components (>1000 different types), and its unrivaled capacity to dynamically transform in response to specific stimuli or other cellular changes.

No class of Golgi protein is stably associated with the Golgi. Integral membrane proteins associated with the Golgi, including processing enzymes for post-translational protein modification (i.e. mannosidase II, galactosyltransferase, etc), are continuously exiting and re-entering the Golgi by membrane trafficking pathways leading to and from the ER. Peripheral membrane proteins associated with the Golgi exchange constantly between membrane and cytosolic pools. Newly synthesized cargo proteins passing through the Golgi to other destinations (which include both integral membrane and luminal proteins) also spend relatively short periods of time in the Golgi. The residency times for these different classes of Golgi proteins vary enormously: Golgi processing enzymes stay for ~60 min, cargo proteins ~30 min, cargo receptors ~10 min and peripheral proteins ~1 min.

 

 

 

 

 

Classification of proteins identified in a rat liver Golgi proteome: (a) Proteins were grouped by cellular location. Twenty-six percent of the identified proteins were known Golgi proteins, whereas 23 percent were ER proteins. Many of the cytosolic and cytoskeletal proteins functionally interact with the Golgi; PM: plasma membrane, Unk: unknown. (b) Golgi-localized proteins were grouped by function.

 

Post-translational protein modifications in the Golgi complex   

The Golgi complex is compartmentalized. Phosphorylation occurs in the cis region. In other regions, different types of carbohydrates are added as a glycoprotein passes through the cisternae. The figure on the right illustrates the different regions where sugars like mannose (man), galactose (gal), etc are added. The final sorting is done in the trans-Golgi complex. Proteins of all living organisms are generally modified in many different ways. A functionally important posttranslational modification is the phosphorylation of proteins. The presence or absence of a phosphate group at specific hydroxyl amino acids regulates the activity, stability, localization, and oligomerization of proteins and in this way influences the flow through metabolic pathways, the transduction of external and internal signals, as well as the timing of developmental steps. The most complex and at the same time energetically most costly protein modification is, however, the glycosylation of proteins.

 

Types of sugar-peptide bonds in eukaryotes. Asn=asparagine;Ser =serine;Thr =threonine; Hyl =hydroxylysine; Hyp =hydroxyproline; Tyr=tyrosine; GlcNAc =N-acetylglucosamine; GalNAc =Nacetylgalactosamine; Glc =glucose; Gal =galactose; Rha =rhamnose; Xyl=xylose; Ara =arabinose; Man =mannose; Fuc=fucose.

 

Processing and maturation of N-glycan chains

Glycoproteins are proteins containing covalently linked oligosaccharides that consist of different monomers and are mostly branched. The carbohydrate moiety amounts to about 20% of the molecular weight, but can be as much as 90% in some cases. In animal cells, glycoproteins are distinguished from proteoglycans, extracellular proteins with mostly long, unbranched polysaccharides, consisting of serially repeating units. The so-called N-glycosylated proteins contain oligosaccharides that are N-glycosidically linked to the g-amido group of asparagines. This type of glycoprotein has been most intensively studied with respect to their structure, biosynthesis, and function.

After the first trimming steps in the ER, the exit of the correctly folded glycoprotein (symbolized by a green ellipse, Figure below) occurs from the ER to the Golgi apparatus where, in a strictly defined reaction sequence, a further demannosylation takes place, followed by transfer of a GlcNAc residue, and finally removal of two further mannose residues. In terminal glycosylation reactions, the mature glycan structure is built up in a protein-dependent, tissue- and organism-specific manner. The generated glycans are classified as high-mannose-type, complex- type, and hybrid-type glycan structures. In the Figure below, only one possible terminal pathway is depicted leading to a biantennary-complex-type glycan chain; the number of antennae may vary up to six. In the case of soluble, lysosomal glycoproteins, a mannose-6-phosphate determinant is generated that functions as the signal for targeting the protein to the lysosome.

The many different sugars, which are either N- or O-glycosidically linked to the amino acid asparagine or to the hydroxy amino acids threonine, serine, hydroxyproline, hydroxylysine, and tyrosine, reflect the complexity of protein glycosylation. Protein N-glycosylation and protein O-mannosylation are evolutionarily conserved from yeast to man. In addition, there is a large variety of more or less highly branched oligo- and polysaccharides of varying composition that are linked to the proximal sugar of the protein. The functional importance of protein glycosylation, however, remained poorly understood for a long time, apart from the role of saccharides as blood-group antigens. Only within the last few years has it become increasingly evident that the lack of individual glycosyl transferases contributing to the synthesis of sugar “trees” of specific proteins can cause most severe congenital diseases in children, including the CDG syndrome (congenital disorders of glycosylation) as well as congenital muscular dystrophies with neuronal-cell-migration defects. Although the molecular details leading to these diseases are only vaguely understood, it seems clear that sugar components of proteins play a major role in embryonic and postembryonic development of humans as well as of all higher eukaryotes.

 

 

Glycoproteins processed by the Golgi apparatus, illustrating species variation.

 

See the movies "
Protein trafficking in the Golgi", "Protein modifications in the Golgi", "Constitutive secretion" and "Regulated secretion" (use Quicktime).

 

Golgi maintenance and biogenesis

The Golgi apparatus contains thousands of different types of integral and peripheral membrane proteins, perhaps more than any other intracellular organelle. To understand these proteins' roles in Golgi function and in broader cellular processes, it is useful to categorize them according to their contribution to Golgi creation and maintenance. This is because all of the Golgi's functions derive from its ability to maintain steady-state pools of particular proteins and lipids, which in turn relies on the Golgi's dynamic character - that is, its ongoing state of transformation and outgrowth from the ER.

The Golgi apparatus performs three major functions essential for growth, homeostasis and division of eukaryotic cells. First, it operates as a carbohydrate factory for the processing and modification of proteins and lipids moving through the secretory pathway. Second, it serves as a station for protein sorting and transport, receiving membrane from the ER and delivering it to the plasma membrane or other intracellular sites. Finally, it acts as a membrane scaffold onto which diverse signaling, sorting and cytoskeleton proteins adhere.

These distinct Golgi functions operate within a structure that is unique among subcellular organelles in many ways, including its composition as a stacked array of cisternae and connecting tubules/vesicles, its enormous diversity of protein components (>1000 different types), and its unrivaled capacity to dynamically transform in response to specific stimuli or other cellular changes. Examples of the Golgi’s dynamic behavior include its reversible disassembly during mitosis and under experimentally induced conditions (e.g. osmotic stress or treatment with BFA, Exo1 or ilimaquinone), and its rebuilding at peripheral ER export sites in response to microtubule disruption or expression of mutated proteins that function in ER-to-Golgi trafficking.

The Golgi’s ability to transform itself fundamentally under different conditions is probably due to the fact that proteins only associate with it transiently as they move through other pathways in the cell. Conditions that alter the entry or return of these proteins to the Golgi, therefore, will disrupt Golgi structure and function. Also, many proteins associated with the Golgi are part of large protein complexes. Altering the association of one protein in the complex may affect the stability and localization of others, with downstream consequences for Golgi organization and structure.

Integral membrane proteins associated with the Golgi are continuously exiting and re-entering the Golgi by membrane trafficking pathways leading to and from the ER. Peripheral membrane proteins associated with the Golgi, by contrast, exchange constantly between membrane and cytosolic pools. Newly synthesized cargo proteins passing through the Golgi to other destinations (which include both integral membrane and luminal proteins) also spend relatively short periods of time in the Golgi. The residency times for these different classes of Golgi proteins vary enormously: Golgi processing enzymes stay for ~60 min, cargo proteins ~30 min, cargo receptors ~10 min and peripheral proteins ~1 min.

 

Post-translational modifications of prohormones   

In addition to glycosylation and phosphorylation in the early stages of the secretory pathway, a secretory protein can undergo sulfation at carbohydrate side chains or tyrosine residues in the late pathway (by carbohydrate and tyrosylprotein sulfotransferases, respectively). Furthermore, the protein can undergo proteolytic cleavage. The first step in proprotein proteolytic processing is usually an endoproteolytic cleavage in the trans-Golgi network (TGN) or immature secretory granule (ISG) on the carboxy-terminal side of a recognition site, often a pair of basic amino acids and mostly Lys-Arg or Arg-Arg. The proprotein processing enzymes are calcium- and pH-dependent serine endoproteases related to the bacterial proteolytic enzyme subtilisin and are called proprotein convertases (PCs). These enzymes are structurally related (strongest in their active sites) and form a gene family consisting of at least seven members, including furin (also called PACE, Paired basic Amino acid residue Cleaving Enzyme), the prohormone convertases PC1/PC3 and PC2, and (in yeast) KEX2. Furin and KEX2 are active in constitutively secreting cells. The neuroendocrine-specific enzymes PC1/PC3 and PC2 cleave prohormones and are selectively present in cells equipped with the regulated secretory pathway. A differential expression of PC1/PC3 and PC2 may result in a different secretory output from endocrine and neuronal cells. Furthermore, a single prohormone may give rise to multiple peptide hormones with a variety of bioactivities. For example, processing of proopiomelanocortin (POMC) can result in a-, b- and g-melanocyte-stimulating hormones (MSHs), the stress hormone adrenocorticotropin (ACTH) and the endorphins, peptides with endogenous opiate-like activity. POMC cleavage by the PC1/PC3 enzyme generates ACTH, whereas PC2 produces the MSHs and endorphins. PCs become active by autocatalytic cleavage of an amino-terminal propeptide that may act as an intramolecular chaperone for proenzyme folding. The 7B2 protein has been found to act as a chaperone specific for PC2 by transiently interacting with the proenzyme form, facilitating proPC2 transport and activation. The proSAAS protein, with a structural organization similar to 7B2, appears to have PC1/PC3 as its major intracellular binding target. The proper acidic environment in the subcompartments of the secretory pathway, essential for optimal PC cleavage activity, is supplied by a H+-pumping vacuolar-type ATPase (V-ATPase). Following cleavage by PCs, exoproteolytic removal of the exposed carboxy-terminal basic residues occurs by the enzyme car­boxypep­tidase E (or KEX1 in yeast). Finally, the generated peptide may undergo one or two modifications that are crucial for its biological activity, namely acetylation at the amino terminus and amidation at the carboxy terminus if the peptide ends in glycine (by the enzyme Peptidyl-glycine-α-Amidating Monooxygenase, PAM).

 

 

 

 

 

 

Proteolytic processing of the prohormones proinsulin and proopiomelanocortin (POMC)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Exocytosis and endocytosis   

Secretory proteins are released into the extracellular space by exocytosis, a process involving the fusion of the secretory granule membrane with the cell membrane. For the two types of exocytosis, namely regulated (triggered) and constitutive (untriggered) exocytosis, regulatory mechanisms of membrane fusion might be quite different. Regulated exocytosis is coupled in most cases with endocytosis to provide a membrane shuttle in order to prevent that the plasma membrane would become too large which would occur when only exocytotic events take place. A number of components of the exocytotic and endocytotic machinery has been recently isolated and identified, e.g. synaptophysin, synaptotagmin, synaptobrevin (VAMP), syntaxin and annexins.

Exocytosis    

Exocytosis is thus the process whereby intracellular fluid-filled vesicles fuse with the plasma membrane (PM), incorporating vesicle proteins and lipids into the plasma membrane and releasing vesicle contents into the extracellular milieu. This membrane fusion event thus mediates the targeting of proteins and lipids to the PM and the secretion of molecules from the cell. Exocytosis can occur constitutively or can be tightly regulated, for example, neurotransmitter release from nerve endings. The last two decades have witnessed the identification of a vast array of proteins and protein complexes essential for exocytosis. As mentioned above, SNARE proteins are probable mediators of membrane fusion, whereas other proteins function as essential SNARE regulators. A central question that remains unanswered is how exocytic proteins and protein complexes are spatially regulated.

Constitutive exocytosis events include the fusion of vesicles derived from the trans-Golgi network (TGN) with the PM, which is essential for the insertion of newly synthesized proteins and lipids into the PM. Polarized cells have developed specialized mechanisms for the targeting of these TGN-derived vesicles to specific regions of the PM, for example, apical vs. basolateral membrane in polarized epithelial cells. In addition, proteins that are constitutively recycled through the endosomal system, such as the transferrin receptor, are transported to the cell surface via the fusion of endosomal vesicles with the PM. These constitutive pathways operate in all cells. In addition, a number of cell types undergo a more specialized form of exocytosis known as regulated exocytosis. Exocytosis of regulated secretory vesicles only occurs upon receipt of a specific stimulus, such as exocytosis of synaptic vesicles in nerve cells. In the majority of cases, regulated exocytosis is stimulated by a local and transient increase in calcium levels.

Exocytosis can involve the full fusion of a vesicle with the PM or, in more specialized cases such as regulated exocytosis from neuronal and neuroendocrine cells, can also occur by a "kiss-and-run" mechanism. Kiss-and-run exocytosis involves the formation of a transient fusion pore that allows release of a limited amount of the vesicle content before the pore re-seals and the vesicle is released from the plasma membrane.

The study of regulated exocytosis at the molecular level has been driven by a number of key questions: How are vesicles recruited to the PM? What proteins anchor vesicles to the PM? What proteins sense and respond to elevated calcium levels? What proteins initiate and catalyze membrane fusion? Finally, how are all these proteins regulated? These questions have led to the identification of numerous proteins, each with their own intricate contribution to exocytosis (see figure). The high efficiency of the exocytosis machinery is exemplified in the ultra-fast response it can exhibit to appropriate stimuli; for example, synaptic vesicles can fuse with the presynaptic PM microseconds following calcium influx.

Lipid rafts: membrane platforms regulating intracellular pathways

Recent studies suggest that lipid rafts, cholesterol and sphingolipid-rich microdomains, enriched in the plasma membrane, play an essential role in regulated exocytosis pathways. The association of SNAREs with lipid rafts acts to concentrate these proteins at defined sites of the plasma membrane. Furthermore, cholesterol depletion inhibits regulated exocytosis, suggesting that lipid raft domains play a key role in the regulation of exocytosis. The role of lipid rafts in regulated exocytosis can vary from a passive role as spatial coordinator of exocytic proteins to a direct role in the membrane fusion reaction.

The lipids that compose cellular membranes are diverse, and as such have different affinities towards proteins and other lipids. The lipid "raft" hypothesis suggests that sphingolipids and cholesterol cluster into discrete regions of the cell membrane. These sphingolipid- and cholesterol-rich domains have been termed lipid "rafts" because they exist in a less fluid and more ordered state than glycerophospholipid-rich domains of the membrane. Studies on model membranes clearly demonstrate clustering and segregation of sphingolipids, cholesterol and certain types of glycerophospholipids, and there is now also evidence that raft-type domains exist in living cells.

Lipid rafts are resistant to solubilization by cold nonionic detergents; this resistance has been used as the criterion for raft purification from numerous cell types, and has allowed a detailed analysis of raft function in various cellular pathways. The term "lipid raft" thus refers to membrane domains that are resistant to detergent extraction. In addition to the characterization of detergent-insoluble lipid rafts, fluorescent imaging techniques such as fluorescence energy transfer and patching/co-patching of membrane proteins have provided essential data on the domain structure of the plasma membrane in fixed and living cells.

The ability of lipid rafts to sequester specific proteins and to exclude others makes them ideally suited to spatially organize cellular pathways. Rafts have been implicated in the regulation of a range of signal transduction pathways, where raft association of components of a signaling cascade likely facilitates protein-protein interactions and signal amplification. Rafts have also been implicated in membrane traffic pathways, such as the formation of regulated and constitutive secretory vesicles.

Evidence that lipid rafts play a key role in regulated exocytosis has emerged from a number of recent studies examining the membrane domain distribution of SNARE proteins. The conclusions form these studies: SNARE proteins are clustered in the plasma membrane in a cholesterol-dependent manner; a proportion of these cholesterol-rich clusters correspond to lipid raft domains - the amount of SNAREs in lipid raft domains depends upon the specific SNARE isoform and the cell type; the integrity of lipid raft domains is important for exocytosis.

Rafts and the regulation of membrane fusion and exocytosis

If SNAREs are localized in different domains at the plasma membrane, what does this mean for membrane fusion? Specifically, do the SNARE clusters in lipid rafts or nonrafts mark the sites of exocytosis? Further analyses will hopefully address whether regulated exocytosis occurs at specific SNARE clusters and determine the molecular characteristics of these fusion-competent domains.

The different protein and lipid composition of lipid rafts and nonraft domains is likely to impact directly on membrane fusion and exocytosis. There are a number of possibilities for how lipid rafts may function in regulated exocytosis (Figure 5): The protein/lipid composition of rafts is conducive to exocytosis, whereas membrane fusion in nonraft domains is prevented (Figure 5A). The figure depicts fusion in the middle of the lipid raft domain; alternatively, membrane fusion may occur at the edges of raft domains, or rafts may surround the fusion site. Fusion occurs exclusively in nonraft domains (Figure 5B). Lipid rafts and nonraft domains support different types of exocytosis, for example full fusion vs. kiss-and-run exocytosis. In the latter form of exocytosis, a transient fusion pore is formed, allowing release of vesicle contents; this fusion pore then quickly seals and the vesicle is released from the plasma membrane. The lipid composition of rafts and nonraft domains may favor one of these exocytic events (Figure 5C,D). Raft and nonraft domains may also regulate the fusion of specific vesicle pools (Figure 5E)

   

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Figure 5. Possible models for the spatial segregation of exocytic events in raft and nonraft domains of the plasma membrane. A) Rafts are the preferred site for all exocytic events. Fusion may occur within a raft domain as shown or alternatively, rafts may surround the fusion site. B) Exocytosis predominates in nonraft domains. C) Lipid rafts mark the site for kiss-and-run and similar transient exocytic events, whereas full fusion of the vesicle and plasma membrane occurs in nonraft domains. D) Kiss-and-run and similar types of exocytosis occur predominantly in nonraft domains, whereas full fusion of the vesicle and plasma membrane occurs in lipid rafts. E) Raft and nonraft domains may regulate the docking and fusion of specific pools of vesicles (the Figure shows one vesicle pool with blue content, the other with green content).

     

 

 

Endocytosis    

Cells interact with their environment at the plasma membrane and through the endocytic pathway. Clathrin-coated pits are specialized regions of the plasma membrane that function to concentrate specific integral membrane receptors. Nutrients, growth factors, viruses and toxins, and immunoglobulins are amoung the many ligands known to bind with high affinity to receptors on the cell surface. The clathrin coated-pit regions of membrane pinch off into the cell, delivering vesicles concentrated with receptors and ligands. This complex, multistep process, known as receptor-mediated endocytosis, involves numerous structural proteins and accessory factors. The figure diagrams the process of endocytic uptake of receptors into clathrin-coated vesicles at the cell surface and the subsequent recycling of involved proteins for further rounds of uptake.

 

 

 

Endocrine diseases linked to the secretory pathway    

A detailed understanding of the processes by which peptide hormones are produced and released, is desirable as a growing number of endocrine disorders is linked to malfunctioning transport, sorting and processing mechanisms in the secretory pathway. For instance, mutations in the thyroid prohormone thyroglobulin may lead to defective prohormone folding and assembly, reduced ER-export and an ER storage disease (hypothyroidism); see below for more examples. In patiens with familial hyperproinsulinemia, a mutation in the B-chain of proinsulin probably causes improper prohormone folding and sorting with diversion of the mutant molecule from the regulated to the constitutive pathway. Alternatively, in such patiens an altered prohormone processing site in proinsulin may occur that results in partially cleaved proinsulin intermediates. Familial neurohypophyseal diabetes insipidus may be caused by mutations in the vasopressin-neurophysin II prohormone gene leading to inefficient signal peptide cleavage, impaired ER-export or improper prohormone folding and resulting in reduced vasopressin production. Familial hypoparathyroidism is due to a mutation in the gene region encoding the signal peptide of preproparathyroid hormone. A mutated exon-intron splice junction gives rise to exon skipping during growth hormone gene expression, resulting in misfolding of the mutant growth hormone that causes Golgi fragmentation, disrupted ER-to-Golgi traffic and familial growth hormone deficiency. Another growth disorder (Laron-type dwarfism) can be caused by a mutation in the growth hormone receptor gene, leading to defective membrane expression of the receptor. A form of diabetes mellitus appears to be due to a mutant insulin receptor leading to disrupted receptor folding and transport. Finally, PC1/PC3 inactivating gene mutations lead to multiple endocrine deficits resulting from defective prohormone processing (obesity, hyperproinsulinemia, hypogonadism).

 

 


Next page: Receptor mechanisms Go back to:  Protein degradation in the cell