Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Cold Spring Harb Perspect Biol. 2011 Nov; 3(11): a005231.
PMCID: PMC3220356
PMID: 21844168

COPI Budding within the Golgi Stack

Abstract

The Golgi serves as a hub for intracellular membrane traffic in the eukaryotic cell. Transport within the early secretory pathway, that is within the Golgi and from the Golgi to the endoplasmic reticulum, is mediated by COPI-coated vesicles. The COPI coat shares structural features with the clathrin coat, but differs in the mechanisms of cargo sorting and vesicle formation. The small GTPase Arf1 initiates coating on activation and recruits en bloc the stable heptameric protein complex coatomer that resembles the inner and the outer shells of clathrin-coated vesicles. Different binding sites exist in coatomer for membrane machinery and for the sorting of various classes of cargo proteins. During the budding of a COPI vesicle, lipids are sorted to give a liquid-disordered phase composition. For the release of a COPI-coated vesicle, coatomer and Arf cooperate to mediate membrane separation.

COPI-coated vesicles transport cargo within the Golgi and from the Golgi to the ER. Arf1 initiates coating and recruits coatomer; after vesicle budding, Arf1 and coatomer cooperate to mediate vesicle release.

Eukaryotic cells are organized as a collection of spatially separated internal organelles embedded in the cytoplasm. Communication between these internal compartments is mediated by trafficking events, some of which are accomplished by vesicular transport. Lipids and proteins are sorted at the membrane of a donor compartment and included into spherical transport carriers with a typical diameter of 50–100 nm. After detachment, these carriers travel through the cytoplasm, and fuse with their target compartment to deliver their content. Trafficking can serve the targeted delivery of newly synthesized proteins and lipids, the uptake of extracellular cargo, and regulatory processes. The Golgi represents a cellular trafficking hub strategically positioned between the endoplasmic reticulum (ER), the site of synthesis of secretory and membrane proteins, and endocytic compartments. Its characteristic stack of cisternae combines the early secretory pathway receiving cargo from the ER via the intermediate compartment (IC, tubulo-vesicular structures between ER and Golgi) with the endosomal membrane system and the plasma membrane. The process of vesicle budding is mainly controlled by vesicular coat proteins, protein complexes able to simultaneously select cargo to be included in a transport carrier and to deform a membrane. Table 1 summarizes the various coats identified to date, as well as their sites of action. Although COPII contributes to the import of newly synthesized proteins to the cis-Golgi (Lee et al. 2004), and retromer mediates input from the endosomes to the trans-Golgi (Bonifacino and Hurley 2008), the clathrin coat (McMahon and Mills 2004) and the recently discovered exomer (Wang et al. 2006) participate in the exit of molecules from the trans-Golgi network (TGN). Coatomer, the coat complex of COPI-coated vesicles, plays various roles at the interface between the ER and the Golgi, as well as within the Golgi (Beck et al. 2009b).

Table 1.

Vesicular coats identified to date, their subunit compositions, sites of action, and roles

Coat/AdaptorStructureSite of actionRoleReferences
ClathrinClathrin Heavy Chain
Clathrin Light Chain
TGN, endosomes, plasma membraneEndocytosis TGN endosome sorting Early-late endosome sortingPearse 1976; McMahon and Mills 2004
AP-1γβ1σ1 µ1TGN, endosomesTGN-endosome sortingPearse and Robinson 1984; McMahon and Mills 2004
AP-2αβ2σ2 µ2Plasma membranePlasma membrane endocytosisPearse and Robinson 1984; McMahon and Mills 2004
AP-3δβ3σ3 µ3EndosomesMelanosome biogenesisMurphy et al. 1991; McMahon and Mills 2004
AP-4εβ4σ4 µ4TGNBasolateral sorting TGN-endosome sortingDell’Angelica et al. 1999; McMahon and Mills 2004
GGA1-3GGA1-3TGNTGN-endosome/-lysosome sortingHirst et al. 2000; Takatsu et al. 2000; McMahon and Mills 2004
COPIαβ′ε/βγδζER, Golgi, Intermediate CompartmentSorting at the ER-Golgi interface and within the Golgi Endosomal functions (see text)Duden et al. 1991; Serafini et al. 1991b; Waters et al. 1991; Harrison-Lavoie et al. 1993; Stenbeck et al. 1993 Bethune et al. 2006
COPIISec13,31/ Sec23,24ERProtein export from the ERBarlowe et al. 1994; Hughes and Stephens 2008
ESCRTHrs-STAM1-2 Vps23,28,37-MvB12 Vps22,25,36 Vps20,24,2-Snf7EndosomesMultivesicular body formation (lysosomal pathway) Cytokinesis AutophagyKatzmann et al. 2001; Hurley and Hanson 2010
RetromerSNX1,2,5,6 Vps26,29-35Early endosomeEndosome-TGN sortingSeaman et al. 1998; Bonifacino and Hurley 2008
ExomeraChs5,6-Bch1,2-Bud7Golgi/endosomeTGN-plasma membrane sortingWang et al. 2006
BBsomeBBS1,2,4,5,7,8,9,10Primary ciliumPlasma membrane-ciliary membrane sortingNachury et al. 2007; Nachury et al. 2010

Coats involved in trafficking steps from or to the Golgi are highlighted in bold.

aExomer is restricted to yeast and fungi that expressed chitin.

ROLES OF THE COPI COAT

Role in the Early Secretory Pathway

Coatomer is a stable protein complex of seven subunits. According to immunogold electron microscopy (EM) studies, membrane associated coatomer is mainly localized to the Golgi apparatus (Duden et al. 1991; Serafini et al. 1991b; Griffiths et al. 1995), IC (Griffiths et al. 1995), coated vesicles surrounding the Golgi (COPI vesicles) (Duden et al. 1991; Serafini et al. 1991b; Griffiths et al. 1995), and the ER (Orci et al. 1994). These locations highlight a role of coatomer in transport at the ER-Golgi interface. COPI vesicles have been shown to participate in the retrieval of proteins from the Golgi back to the ER (Cosson and Letourneur 1994; Letourneur et al. 1994). In contrast, a role of these carriers in anterograde ER-to-Golgi transport, is still a matter of debate. Newly synthesized proteins are first exported from the ER to the IC by COPII vesicles (Aridor et al. 1995; Scales et al. 1997). Inhibition of COPI transport by a dominant negative mutant of Arf1, microinjection of anti-ß-COP antibodies, or BFA results in an impairment of subsequent IC-to-Golgi trafficking, which can be interpreted either as a direct effect, or as a secondary consequence of the default of coatomer-mediated recycling of membrane components of the COPII system (for a review see Pelham 1994), and/or as a fault in IC maturation (Aridor et al. 1995; Scales et al. 1997; Shima et al. 1999).

The prominent presence of COPI vesicles around the cisternae of the Golgi suggests a role of these carriers in intra-Golgi trafficking (Duden et al. 1991; Serafini et al. 1991b). Although these carriers have been known for more than 20 years, their roles are far from being completely established, and several models exist with regard to their functions (see other chapters of this book). According to the vesicular model, the Golgi is static, and anterograde COPI vesicles deliver newly synthesized proteins and lipids to successive cisternae. In the maturation and progression model, cisternae form at the cis-Golgi and progress along the Golgi to finally disassemble at the trans-Golgi. Here retrograde transport of Golgi resident proteins by COPI vesicles would ensure the maturation of the cisternae along the cis-trans axis of the Golgi. Observation of both anterograde and retrograde cargoes within COPI vesicles (Orci et al. 1997) gave rise to the percolating model, a combination of the two mechanisms above (Orci et al. 2000). COPI vesicles would mediate bidirectional transport between two adjacent cisternae by a random walk. Anterograde transport is simply driven by protein biosynthesis, entry of biosynthetic cargos at the cis-Golgi and their exit at the trans-Golgi, generating a flow across the Golgi.

Lipids droplets, another kind of structure originating at the ER-Golgi interface, also require coatomer activity for proper functioning (Guo et al. 2008). By mediating the delivery of enzymes to this compartment, the COPI system might help control the homeostasis of lipids in the cell (Beller et al. 2008; Soni et al. 2009), although carriers with cargo indicative of this pathway have not yet been described.

Role in Mitosis and Golgi Positioning

In addition to its role in interphase, coatomer has an active role during mitosis. COPI activity, concomitantly with repression of COPII vesicle formation (Farmaki et al. 1999), strongly contributes to the fragmentation of the Golgi apparatus into vesicles (Misteli and Warren 1994; Tang et al. 2008). In addition, recruitment of coatomer by a nucleoporin may induce the breakdown of the nuclear envelope (Liu et al. 2003).

An interaction of coatomer with cdc42 (Wu et al. 2000) and dynein was attributed to positioning of the Golgi (Chen et al. 2005; Hehnly et al. 2010).

Role in the Endocytic Pathway?

The early secretory and the endocytic pathways are similar and mirror each other as newly synthesized protein from the ER and proteins of the plasma membrane follow a similar trafficking scheme. Both protein populations are first transferred to a sorting station, the Golgi apparatus or the early endosomes, before being either recycled back to their starting compartment (ER or plasma membrane), or further transported along the late secretory pathway or to the late endosomal or lysosomal compartments. Pools of coatomer have also been identified at endosomal membranes (Whitney et al. 1995; Aniento et al. 1996; Gu and Gruenberg 2000). In this context it is of note that coatomer has been implicated to take part in the maturation of early endosomes (Whitney et al. 1995; Aniento et al. 1996; Daro et al. 1997; Gu et al. 1997; Gu and Gruenberg 2000; Gabriely et al. 2007), and/or recycling toward the plasma membrane (Daro et al. 1997; Razi et al. 2009). Moreover, coatomer may participate in the maturation of specialized endosomes such as phagosomes (Botelho et al. 2000; Beron et al. 2001; Hackam et al. 2001), autophagosomes (Razi et al. 2009), and peroxisomes (Lay et al. 2006). It is, however, not known if in these endosomal functions coatomer directly serves as a coat, or if these processes directly depend on the service of COPI-coated vesicles. Furthermore, recruitment of coatomer to endosomes, in contrast to Golgi binding, was reported to be pH-sensitive (Aniento et al. 1996), and to require only a subset of coatomer subunits (Whitney et al. 1995; Aniento et al. 1996), suggesting differences in mechanisms underlying the function of the complex in the endocytic and in the early secretory pathways.

How Can COPI Mediate Various Trafficking Steps?

How coatomer could serve different trafficking routes does not seem trivial to explain at first sight, specifically because only one form of the complex was known to begin with. The complex is organized in two subcomplexes: a trimer composed of α-COP, β′ε- COP, and ε-COP, and a tetramer of β-COP, γ-COP, δ-COP, and ζ- COP (Lowe and Kreis 1995; Fiedler et al. 1996a; Pavel et al. 1998). More recently, the two coatomer subunits γ-COP and ζ- COP were found to exist in two isoforms, called γ1, γ2, and ζ1, ζ2, (Futatsumori et al. 2000). Each isoform is, like all other subunits, present in coatomer as one copy, resulting in four possible different heptameric protein complexes (Wegmann et al. 2004). These coatomer isoforms localize differently within the Golgi apparatus of mammalian cells (Moelleken et al. 2007), suggesting different sites of budding for each of them. COPI shares structural aspects and mechanistic properties with the well-characterized clathrin coat (see Fig. 1). Clathrin heavy and light chains polymerize into a cage constituting the outer layer of the coat. An inner layer composed of adaptor proteins takes the role of a link between the outer layer and the membrane. Various adaptors allow recruitment of the clathrin system to different organelle membranes, and thus selection of different sets of cargo for inclusion into coated vesicles (Robinson 2004).

An external file that holds a picture, illustration, etc.
Object name is cshperspect-GOL-005231_F1.jpg

The heptameric complex coatomer is compared with the clathrin/adaptor complex AP2. The stable complex coatomer can be dissociated in vitro into subcomplexes: a trimer (αβ′ε) and a tetramer (βγδζ). Structural similarities exist between the trimeric coatomer subcomplex (Lee and Goldberg 2010) and the triskelion of clathrin heavy and light chains (Xing et al. 2010). Within the tetrameric subcomplex of coatomer structural similarities are reported for ζ-COP and AP2σ2 (Yu et al. 2009), and for the γ-COP and the β2 appendage domains (Watson et al. 2004). Thus, the trimeric COPI-subcomplex is thought to resemble the clathrin part, and the tetrameric COPI-subcomplex the adaptor part of a clathrin-coated vesicle.

X-ray crystallographic analyses of partial structures revealed similarities between the coatomer and the clathrin coat. On one hand, the three-dimensional structures of coatomer subunits γ- COP and ζ-COP show a striking resemblance to βAP2 (Hoffman et al. 2003; Watson et al. 2004) and σAP2 (Yu et al. 2009), two subunits of the tetrameric clathrin adaptor proteins adaptin (AP2). The overall structure of the tetrameric coatomer subcomplex β/δ/γ/ζ-COP is therefore likely to be similar to the AP complexes. Likewise, the α/β′/ε-COP-trimer presents a spatial organization similar to the clathrin subunits (Hsia and Hoelz 2010; Lee and Goldberg 2010). Thus, it is tempting to speculate that, like the various adaptins, different coatomer isoforms provide ways to modulate the cargo repertoire of the COPI system, leading to distinct pools of coated vesicles involved in different pathways. In agreement with this hypothesis, different subpopulations of COPI vesicles with different cargo compositions can be observed in the living cell (Orci et al. 1997) and in vitro (Lanoix et al. 2001; Malsam et al. 2005 ).

Box.

Lipids and COPI Vesicles

Molecular mechanisms that underlie the sorting of proteins have been elucidated to quite some detail in the recent years. Much less is known, however, of how a living cell can maintain the identities of its various organelles with respect to their unique lipid composition, although lipidomes of a few membrane carriers have become available recently (Brügger et al. 2000; Takamori et al. 2006; Klemm et al. 2009). Lipidomic analysis of COPI vesicles revealed a depletion of cholesterol and sphingomyelin (SM), two lipids characteristic of the liquid-ordered (Lo) phase, when compared with the donor Golgi membrane (Brügger et al. 2000). In vitro analysis of vesicle formation from giant unilamellar vesicles supports the idea that COPI budding occurs exclusively from liquid-disordered (Ld) phases (Manneville et al. 2008). In accordance, ER membranes contain less cholesterol and SM than Golgi membranes. Combined with retrograde transport by COPI vesicles from the Golgi to the ER, selective transport of Ld lipids would allow maintaining lipid homeostasis within the early secretory pathway. In addition, selective sorting of Ld lipids by the COPI coat could induce a phase separation within the forming bud. The line tension thus created between Ld and Lo phase would favor fission of the formed vesicle (for a review, see Pinot et al. 2010).

Additionally, an active role of lipids in COPI vesicle formation has been proposed. Lipids like phosphatidic acid (PA) (Yang et al. 2008), diacylglycerol (DAG) (Fernandez-Ulibarri et al. 2007; Asp et al. 2009), and phosphatidyinositol (PI) (Simon et al. 1998; Carvou et al. 2010) can affect membrane curvature and thus have the potential of facilitating budding and/or scission of vesicles. However, a direct involvement in budding/fission of lipids is experimentally not easy to access, and data existing so far are indirect observations based on inhibition of lipid modifying enzymes. Furthermore, none of these enzymes seems to be required for vesicle formation in in vitro reconstitution experiments. More investigation is thus needed to clarify whether local regulation of membrane lipid composition is a basis for fine-tuning of the formation of a COPI vesicle in the living cell.

MOLECULAR MECHANISM OF COPI VESICLE FORMATION

The generation of a fusion-competent COPI transport vesicle can be conceptually subdivided into partially interdependent steps: coat recruitment, uptake of cargo, budding, membrane separation (scission), and uncoating (see Fig. 2). Whereas the minimal requirements for coat formation, budding, fission, and uncoating have been defined by reconstitution experiments using chemically defined lipids and purified protein components (Spang et al. 1998; Bremser et al. 1999; Reinhard et al. 1999), the mechanisms underlying cargo uptake are still not completely understood.

An external file that holds a picture, illustration, etc.
Object name is cshperspect-GOL-005231_F2.jpg

Individual steps in the formation of a COPI vesicle. (Scheme adapted from Beck et al. [2009b] and reprinted, with permission, from Elsevier © 2009.) For details see text.

Coat Recruitment

The small Ras-like GTPase ADP-ribosylation factor 1 (Arf1) plays a central role in the formation of COPI-coated vesicles at the Golgi apparatus. However, action of Arf1 is not limited to this organelle and can function at various intracellular compartments, involving different sets of effectors (Nie et al. 2003; Donaldson et al. 2005; Volpicelli-Daley et al. 2005). It is thus very likely that additional factors are required to provide a spatio-temporal regulation of coatomer recruitment at specific sites suitable for COPI vesicle biogenesis. Like most of the small GTPases, Arf1 cycles between a GDP-loaded inactive cytosolic state and an active, membrane anchored, GTP-loaded state (Randazzo et al. 1993). This molecular switch is positively regulated by guanine nucleotide exchange factors (GEFs), which catalyze exchange of GDP with GTP. On activation of an intrinsic low GTPase activity by Arf GTPase activating proteins (ArfGAPs), Arf1 is converted into its inactive GDP form and released from the membrane.

Specific recruitment of the small GTPase to the membrane is a prerequisite for its function. In agreement with its role in COPI vesicle formation, Arf1 contains Golgi localization signals encoded in its sequence. Arf1-GDP binds to a dimeric complex of members of the p24 family, a group of type I transmembrane proteins. In vitro cross-linking experiments reveal a specific interaction between a carboxy-terminal site of Arf1 with a dimer of p23 or p24 (Gommel et al. 2001). Foerster resonance energy transfer (FRET) experiments confirmed the existence of this interaction in the living cell (Majoul et al. 2001). In addition, at the cis-Golgi, Arf1 is recruited to the membrane via membrin, a SNARE protein. This interaction is mediated by a MXXE motif within the GTPase. Mutation of this motif impaired recruitment of Arf1 to the cis-Golgi without affecting its binding to the trans-Golgi (Honda et al. 2005). Thus, other signals within Arf1 must mediate targeting of the GTPase to the trans-Golgi.

Arf1 is activated on membranes by large ArfGEFs that are members of the Sec7 super family. Based on sequence homologies, mammalian ArfGEFs can be classified into the following groups: (1) Golgi Brefeldin A (BFA)-resistance factor 1/BFA-inhibited GEF (GBF/BIG), (2) Arf nucleotide binding site opener (ARNO)/cytohesins, (3) exchange factor for Arf6 (EFA6), (4) BFA-resistant Arf GEF (BRAG), and (5) F-box only protein 8 (FBX8) (Casanova 2007).

GBF1 and the BIGs, as well as their yeast homologs (Gea1/2 and Sec7p, respectively) activate Arf1 at the Golgi membrane (Claude et al. 1999; Spang et al. 2001; Kawamoto et al. 2002; Zhao et al. 2002), whereas the other ArfGEFs act at post-Golgi compartments. The mammalian Golgi GEFs are not evenly distributed across the stack of cisternae. BIG1 and BIG2 are present at the trans-Golgi, the TGN, and the recycling endosomes (Mansour et al. 1999; Shinotsuka et al. 2002; Shin et al. 2004). Their mode of recruitment to the membrane is not well understood. GBF1 localizes to IC and Golgi, acting as a direct mediator of retrograde transport between these compartments and the ER (Kawamoto et al. 2002; Garcia-Mata et al. 2003; Zhao et al. 2006). This localization is directly dependent on an interaction between the GEF and Rab1b (Monetta et al. 2007). Recruitment of GBF1 also requires PI4P, and it was proposed that Rab1 contributes to GBF1 recruitment by locally activating phosphatidylinositol 4-kinase (PI4KIIIalpha) (Dumaresq-Doiron et al. 2010).

Once both Arf1 and its exchange factor are recruited to the membrane, a Sec7 domain within the ArfGEF triggers the activation of the small GTPase. A critical feature of this catalytic domain is its “glutamic finger,” a conserved glutamate residue exposed at the tip of a hydrophilic loop between helices 6 and 7 (Beraud-Dufour et al. 1998). Nucleotide exchange is mediated by electrostatic competition of this glutamate side chain with the nucleotide’s β–phosphate. On exchange of GDP to GTP, a conformational change within Arf1 leads to exposure of its amino-terminal amphipathic and myristoylated helix, which in turn causes insertion of this helix into the lipid bilayer and thereby secures membrane anchorage of Arf1 (Franco et al. 1996; Antonny et al. 1997). This model was challenged by recent structural data. Nuclear magnetic resonance spectrometry of full-length myristoylated Arf1 embedded into bicelles showed an unexpected localization of the myristic acid residue. Instead of being inserted within the bilayer parallel to the phospholipid-acyl chains, the myristic fatty acyl chain appears on top of the bicelles, perpendicular to the phospholipids (Liu et al. 2010). Whether this orientation is because of the physicochemical nature of the bicelles, or reflects Arf1’s positioning on physiological membranes, is presently not known.

Recruitement of coatomer to Golgi membranes is tightly correlated to Arf1 activation (Donaldson et al. 1991; Serafini et al. 1991a; Palmer et al. 1993). The complex is recruited to the Golgi membrane en bloc (Hara-Kuge et al. 1994), in contrast to clathrin/AP and COPII that are composed of two successively recruited layers. Coatomer forms multiple interfaces with Arf1-GTP: site-directed photolabeling studies highlighted specific contacts between the GTPase and the subunits β′-COP, β-COP, δ- COP, as well as the trunk domain of γ-COP (Zhao et al. 1997, 1999; Sun et al. 2007). Yeast two hybrid analysis further points to an interaction of Arf1 with ε-COP (Eugster et al. 2000). Three to four Arf1 molecules bind to one coatomer complex (Serafini et al. 1991a; Beck et al. 2009a). Thus, interaction of coatomer with Arf provides multiple protein–protein interfaces. Binding of γ-COP, via its trunk and appendage domains, to dimers of p24 transmembrane protein family members (COPI transmembrane machinery proteins, see later) further stabilizes coatomer on the Golgi membrane (Harter and Wieland 1998; Bethune et al. 2006). Members of the p24 family carry variants of a dilysine motif in combination with a double F motif, FFXX(K/R)(K/R)Xn (n ≥ 2). Binding of coatomer to these signals depends more strongly on the phenylalanine than on the dilysine residues (Fiedler et al. 1996b; Sohn et al. 1996). These machinery signatures are recognized exclusively by γ-COP (Bethune et al. 2006). p24 proteins are present in COPI vesicles in amounts stoichiometric to coatomer (Sohn et al. 1996). In a yeast strain defective for vesicle fusion, a p24 knockout reduced the number of COPI-coated vesicles (Stamnes et al. 1995). These data suggested that type I transmembrane proteins represent membrane receptors for coatomer and are actively required for the formation of COPI vesicles (Stamnes et al. 1995; Sohn et al. 1996). Surprisingly, however, in yeast the deletion of all p24 proteins showed only a reduction in the rate of transport of some cargo proteins (Springer et al. 2000). More recently, additional biochemical evidence was reported that in yeast the p24 complex participates in retrograde transport from Golgi to ER, by promoting the formation of COPI vesicles (Aguilera-Romero et al. 2008). p24 members are indispensible in mammals: knockout of p23 is lethal at the earliest possible time point in the development of a mouse embryo (Denzel et al. 2000).

Specific binding of the γ-COP trunk domain to dimers of p23 and p24 (not p25, p26, or p27), results in a conformational change in γ-COP (Reinhard et al. 1999; Bethune et al. 2006) that is transmitted to the α-COP subunit (Langer et al. 2008). This spatial rearrangement causes aggregation of the complex and is likely to initiate coatomer polymerization (Reinhard et al. 1999), providing the energy to bend the membrane and sculpting a COPI-coated bud (R Beck et al., unpubl.).

A machinery component of COPI vesicles with another dilysine motif is the seven-helix transmembrane protein KDEL receptor that functions as a transmembrane adaptor. Its cytosolic tail interacts with coatomer via a KKXSXXX signal, active only when its serine residue is phosphorylated (Cabrera et al. 2003), whereas its luminal part interacts with soluble proteins that harbor a C-terminal KDEL-sequence (Lewis and Pelham 1992). As a result, KDEL-proteins are included into COPI vesicles and retrieved to the ER (Pelham 1991; Majoul et al. 1998), where they dissociate from the KDEL receptor, probably because of a difference of pH between Golgi and ER (Wilson et al. 1993). The free KDEL-receptor is then cycled back to the Golgi.

Identification of a subpopulation of COPI vesicles, which lacks p24 proteins (Malsam et al. 2005), raises the question of whether these carriers contain transmembrane machinery components, other than p24 family proteins, that are crucial for their biogenesis.

Cargo Sorting/Sorting Motifs

Proteins are directed into COPI vesicles by various mechanisms based on direct or indirect binding to the coat.

Membrane proteins to be included into COPI vesicles can also be recognized directly by coatomer through sorting motifs present in their sequence. The first signals characterized were dilysine motifs present at the extreme carboxyl terminus of membrane proteins (Nilsson et al. 1989). They bind directly to coatomer at sites different to the above machinery proteins and induce the retrieval from the Golgi to the ER of the host protein (Cosson and Letourneur 1994; Letourneur et al. 1994). Various subsets of dilysine motifs exist that are recognized differentially by the coatomer complex (Schroder-Kohne et al. 1998). In contrast to the p24 familiy carboxy-terminal signatures, KKXX motifs interact with the WD40 domain of α-COP, and KXKXX binds to a similar domain within β′-COP (Eugster et al. 2004). Thus structural differences constitute the molecular basis for coatomer to discriminate machinery components that are recycled from cargo proteins that are transported and delivered unidirectionally. The affinity between coatomer and such cargo sorting signals depends on the nature of the X amino acids following the lysine residues (Zerangue et al. 2001), thus giving rise to differences in efficiency of retrieval to the ER.

Together with a modulation of the efficiency of ER exit, these tuned affinities may provide a mechanism to control the steady-state localization of membrane proteins at the ER-Golgi interface (Zerangue et al. 2001).

Furthermore, Arginine-based motifs that conform to the consensus sequence (Φ/Ψ/R)RXR (where Φ/Ψ is an aromatic or bulky hydrophobic residue) (Zerangue et al. 1999) are recognized by coatomer subunits β- and δ-COP (Michelsen et al. 2007). In contrast to dilysine motifs, they are not restricted to carboxyl termini and can have a more flexible positioning within the cytosolic tail of the host protein (Shikano and Li 2003). These signals can control the maturation of membrane protein complexes (Michelsen et al. 2005). As long as a complex is not fully assembled, such arginine motifs present in its subunits are exposed, inducing retrograde transport to the Golgi. On complete assembly of a complex, the signals become masked, allowing export of the complex to the cell surface (Zerangue et al. 1999). Several mechanisms for masking of the signals have been proposed, including steric masking by a partner subunit (Michelsen et al. 2005), inactivation of the signal by phosphorylation of nearby residues (Scott et al. 2001), or competition for motif binding of coatomer with 14-3-3 proteins (O’Kelly et al. 2002; Yuan et al. 2003), or PDZ-domain proteins (Standley et al. 2000).

Additional sorting motifs are based on aromatic residues. A “δL” motif confers binding to δ-COP and retrieval to the ER (Cosson et al. 1998), and a FXXXFXXXFXXLL motif in the Dopamin1 receptor mediates interaction with γ-COP, which is necessary for physiological trafficking of the receptor (Bermak et al. 2002). However, not all the membrane proteins transported within COPI vesicles carry coatomer-interacting motifs. Notably, this is the case of glycosylation enzymes with their tails lacking known sorting signal. Recent studies in yeast implicated Vps74 as an essential factor for the packaging of glycosylation enzymes into transport carriers (Schmitz et al. 2008; Tu et al. 2008). By binding simultaneously to coatomer and the cytosolic tails of glycosyltransferases, Vps74 works as a coat-cargo adaptor (Tu et al. 2008). Although further studies will be needed to unravel similar mechanisms in mammalian cells, Vps74 suggests the existence of a set of similar cargo adaptors in higher eukaryotes.

Inclusion of Cargo into COPI Vesicles

Two modes of incorporation of membrane proteins into COPI vesicles are observed. Proteins directly involved in the budding process are incorporated into vesicles in a GTP-independent manner and are further referred to as machinery components (Nickel et al. 1998; Malsam et al. 1999). Cargo proteins, on the other hand, require GTP hydrolysis for their uptake into the transport carrier. Indeed, in the presence of GTPγS, a poorly hydrolysable form of GTP, or Arf1-Q71L (an Arf1 variant locked in its GTP-loaded state), COPI vesicles can still form, but appear devoid of anterograde and retrograde cargo (Nickel et al. 1998; Lanoix et al. 1999; Malsam et al. 1999; Pepperkok et al. 2000). In contrast, uptake of p23, p24, or KDEL receptor is still effective under these conditions (Nickel et al. 1998). As previously described, activation of Arf1 is a prerequisite for the recruitment of coatomer to the membrane. As a corollary, hydrolysis of Arf1-GTP into Arf1-GDP leads to membrane uncoating (see later). This seems in contradiction with a role of GTPase activity in cargo sorting. In in vitro experiments with the COPII coat, priming complexes weakly bound to unassembled cargo are instantaneously released, whereas those interacting more strongly to oligomerized cargo remain transiently associated to the membrane, increasing their probability to be incorporated into a final COPII vesicle (Sato and Nakano 2005). This characteristic may provide the molecular basis of a GTPase-driven kinetic proofreading mechanism (Sato and Nakano 2007; Tabata et al. 2009). Kinetic analysis of the COPI system performed in living cells revealed that Arf1 dissociates faster from membranes than coatomer (Presley et al. 2002), suggesting that coatomer is not immediately released after Arf1 inactivation but stays metastably associated to the membrane for an additional period of time. This difference is likely because of the additional binding of coatomer to membrane proteins (e.g., p24 family proteins), and to lateral interactions of the polymerized complexes within the coat network. It is open at present whether the different dissociation kinetics of Arf and coatomer would reflect a kinetic proofreading mechanism, a model that was suggested by Weiss and Nilsson (2003).

A different, but not mutually exclusive, mechanism has been proposed for specific enrichment of cargo in a nascent COPI vesicle based on the curvature sensitivity of ArfGAP1 (Liu et al. 2005). Preferential binding of the protein to positively curved membranes (Bigay et al. 2003) increases local ArfGAP1 activity. As a result, GTP hydrolysis, and thus coat release, from the positive curvature of a forming bud would be increased. This GTPase activity would lead to a flux of coatomer from the rim, where Arf and coatomer are recruited to the center of a growing bud, thereby mediating cargo concentration (Liu et al. 2005). Several machinery components can locally modulate GTP hydrolysis. Coatomer itself can stimulate ArfGAP1 (Goldberg 1999; Szafer et al. 2001), whereas the cytosolic tails of p23 and p24 slow down the hydrolysis of Arf1 (Goldberg 2000; Lanoix et al. 2001). Interaction between a KDEL receptor and ArfGAP1 may recruit an additional activating protein to the membrane, and thus increase local GTPase activity (Aoe et al. 1997, 1998). This would give rise to another, less characterized, level of regulation.

Budding and Scission

More than 20 years after the discovery of COPI-coated vesicles (Orci et al. 1986; Malhotra et al. 1989), there is still an ongoing discussion about the structural components of this class of vesicular carrier (Beck et al. 2009b; Hsu and Yang 2009; East and Kahn 2010). The two cytosolic components essential for COPI vesicle formation in vitro are Arf1 and the heptameric coat complex coatomer (Serafini et al. 1991a; Orci et al. 1993).

In a chemically defined in vitro system, using synthetic liposomes with a nonphysiological lipid composition, Arf1-GTP and coatomer alone were sufficient to induce vesicle formation (Spang et al. 1998). Nevertheless, in the presence of a cytoplasmic domain of a p24 family protein, vesicle formation is stimulated, and independent of a wide range of lipid compositions (Bremser et al. 1999). As described above, the p24 proteins were initially found to be a major component of COPI-coated vesicles (Stamnes et al. 1995; Sohn et al. 1996) and serve as membrane machinery for these carriers.

More recently a membrane surface activity was observed of the activated, GTP-loaded form of Arf1 on synthetic membranes in the absence of coatomer (Beck et al. 2008; Krauss et al. 2008; Lundmark et al. 2008). This activity results in tubulation of giant unilamelar vesicles (GUVs), or of membrane sheets tethered to glass surfaces, and strictly depends on a dimerization of the small GTPase. This dimerization is essential in the living cell, as a yeast strain devoid of Arfs (Takeuchi et al. 2002) cannot be rescued with an Arf point mutant unable to dimerize (Beck et al. 2008). The surface activity of Arf1 is a prerequisite for COPI vesicle formation (Beck et al. 2008). Thus, during formation of a COPI bud, two different activities have the potential to deform a flat donor membrane into a curved bud. These are the polymerization of coatomer on its recruitment and/or the dimerization-dependent membrane surface activity of activated Arf.

Cryo electron microscopy and tomography of liposomes incubated with Arf wt, GTP and coatomer revealed production of regular free COPI vesicles, as expected from earlier work (Reinhard et al. 2003). With the nondimerizing mutant of Arf, however, hardly any free vesicles are formed. Instead, the liposomes are transformed into flower-like structures that represent multiple buds linked by a continuous bilayer, indicating a block in scission. It is of note that neither mechanical shearing nor high salt conditions are needed to create a free COPI-coated vesicle (R Beck et al., unpubl.). In the mechanistically similar COPII system, a membrane surface activity of Sar1 was revealed earlier (Bielli et al. 2005; Lee et al. 2005). Here, the fission step in the formation of COPII-coated vesicles from liposomes was dependent on the N-terminal amphipathic helix of the small GTPase (Lee et al. 2005). Most recently, a regulated scaffold assembly by Sar1 was observed (Long et al. 2010), opening a possibility that Sar1 assembly on membranes controls constriction of the bud neck similar to the fission mechanism proposed for dynamin in the clathrin system (Bashkirov et al. 2008; Pucadyil and Schmid 2008). However, the dynamin mechanism requires hydrolysis of GTP for the scission process, and in various reports the formation of COPII vesicles was described in the presence of nonhydolyzable analogs of GTP (Barlowe et al. 1994; Oka and Nakano 1994). Likewise, Arf1-dependent formation of COPI vesicles is independent of GTP hydrolysis (Malhotra et al. 1989). Taken together, this indicates that the small GTPases Arf1 and Sar1 not only mediate coat assembly but also promote the fission step to release a nascent vesicle in a mode other than that proposed for dynamin. What mechanism can explain these two functions of a small GTPase?

A Model for Membrane Separation Based on Cooperation of Coatomer with Arf1

Within the growing coat, Arf interacts with the membrane via its myristoylated amphipathic helix, and with its covering layer of coatomer via several defined interfaces (Zhao et al. 1997, 1999; Sun et al. 2007). Two Arf1 dimers are bound to one coatomer complex (Serafini et al. 1991b; Beck et al. 2008) (R Beck et al., unpubl.). In the positive curvature of the growing bud, Arf’s interaction with the membrane is energetically favorable. As the bud becomes more complete, a neck forms at the site of Arf and coatomer recruitment, with increasingly negative curvature, which is highly unfavorable for Arf binding. This energy strain could be relaxed by diffusion of the small GTPase from the negatively curved zone. If the stability of the complex of the dimerized Arfs with the covering network of coatomer is strong enough, however, to prevent Arf from escaping, the local high-energy state can only be relaxed by fusion of the adjacent membranes in the neck, causing membrane separation. The need for Arf to dimerize is simply explained by comparing the stability of a complex of coatomer formed with either two dimers of Arf (Arf wt) or with four monomers (nondimerizing mutant). With Arf wt the complex’s stability is higher because of the energy released by the two Arf dimerization interfaces (R Beck et al., unpubl.).

Additional Proteins Implicated in COPI Fission

A variety of proteins have been shown to be involved in COPI vesicle formation. Among those, Brefeldin-A ADP-ribosylated Substrate (BARS) (Yang et al. 2005), endophilin (Yang et al. 2006), and ArfGAP1 (Yang et al. 2002) are thought to play a role in membrane scission and/or act as a coat component. Some of these functions have been challenged by more recent investigations (Gallop et al. 2005; Beck et al. 2009b).

Whereas the basic mechanisms of the core machinery as deduced from experiments in reconstituted systems seem quite clear, the mechanisms underlying the contributions of such additional components that might play important roles in the living cell represent challenges for future research.

Uncoating

To allow fusion with its target membrane, a COPI-coated vesicle must be uncoated. In a prevailing model, this process is coupled to hydrolysis of GTP by Arf1 (Tanigawa et al. 1993), triggered by ArfGAPs (Cukierman et al. 1995). The members of this protein family are defined by the presence of a catalytically active domain bearing a zinc finger motif, followed by an invariant arginine residue (CX2CX16CX2CX4R) (Cukierman et al. 1995).

Three mammalian ArfGAPs (ArfGAPs1, 2/3) are implicated in the COPI system as knock down of all three ArfGAPs increased the level of membrane bound Arf1 in the living cell. As an additional phenotype, the cycling protein ERGIC-53, the Golgi tethering protein GM130, and coatomer accumulated in the ER-Golgi intermediate compartment. Golgi-to-ER retrograde transport was blocked as consequence of the triple knock down, a phenotype similar to a β-COP knock down (Saitoh et al. 2009). Yeast homologs, Gcs1 for ArfGAP1, and Glo3 for ArfGAPs 2 and 3, function as an essential pair for retrograde transport from the Golgi to the ER (Poon et al. 1999). These data suggest that both types of ArfGAPs in mammals (ArfGAP1 and ArfGAP2/3) and in yeast (Gcs1 and Glo3) are interchangeable.

It came as a surprise when ArfGAP1 activity turned out to be regulated by membrane curvature (Bigay et al. 2003), with increasing enzyme activity correlated to decreasing diameter of a liposome (i.e., increasingly steeper curvature). This sensitivity is based on the presence of a motif termed ALPS (for ArfGAP1 Lipid Packing Sensor) that on membrane contact forms an amphipathic helix that inserts bulky hydrophobic side chains between the loosely packed lipid head-groups of the outer leaflet of a positively curved membrane (Bigay et al. 2005).

More recently comparison of ArfGAP1 with the related ArfGAP2/3 revealed functional differences of these auxiliary proteins. ArfGAPs 2 and 3 lack an ALPS motif and hence do not show sensitivity to membrane curvature (Weimer et al. 2008). In contrast, the activities of ArfGAPs2 and 3 depend on coatomer (Weimer et al. 2008; Kliouchnikov et al. 2009). ArfGAP2 and 3 were shown to directly interact with coatomer (Frigerio et al. 2007; Kliouchnikov et al. 2009), and can also be detected on purified COPI vesicles derived from liposomes or native Golgi membranes (Frigerio et al. 2007; Weimer et al. 2008).

In the living cell ArfGAP2/3 and ArfGAP1 show fundamental differences. ArfGAP2/3 followed the dynamics of membrane association of coatomer more closely than does ArfGAP1. Furthermore, ArfGAP2/3 knock down caused unstacking of the Golgi and cisternal shortening, similar to conditions in which vesicle uncoating was blocked (Kartberg et al. 2010). Taken together, ArfGAP1, independent of coatomer, is likely to drive the cycle of activation and inactivation of Arf1 explained above, whereas the coatomer-dependent ArfGAPs2/3 are candidate-uncoating, GTPase-activating proteins.

A possible contribution to uncoating of tethering proteins emerges from studies in yeast, which requires the Dsl1 complex, consisting of Dsl1p, Dsl3p, and Tip20 (Andag et al. 2001; Tripathi et al. 2009) for Golgi to ER transport. Dsl1p was shown to interact directly with the subunits α and δ of the COPI complex (Reilly et al. 2001; Andag and Schmitt 2003; Zink et al. 2009). The observations that COPI-coated vesicles accumulate in Dsl1p-depleted cells led to the suggestion that this multi protein complex has, with its role in tethering, a function in uncoating (Zink et al. 2009). A more detailed explanation of tethers within the early secretory pathway as well as their functions is given by Malsam and Söllner (2011).

OPEN QUESTIONS

With the X-ray structures available for clathrin and COPII vesicle coat proteins and partial structures of the COPI coat, it becomes clear that common protein modules are used to build the various coats. This structural information relates to crystallized coat protein components, however, the structures of the proteins within the coats they form are still elusive. How do the structures of coat proteins in solution and within the network of polymerized complexes on the coated vesicle compare? This question is of particular interest for COPI vesicles, because in this system coatomer is recruited en bloc, and undergoes a conformational change after its contact with p24 family proteins. Cryo electron microscopic tomography may allow extracting such structural information from coated liposomes or vesicles. Insights derived from a combination of X-ray crystallography and electron microscopy analysis of coats and cocrystals of coat components will then help us understand basic open questions about the molecular mechanism of energy-driven cargo uptake into a vesicle. More biochemical approaches will be needed to elucidate the different roles isoforms of coatomer take in different locations within the Golgi. To this end, coated vesicles need to be prepared with individual coatomer isoforms and their proteomes compared. In this context, it is also of interest to know whether the coatomer-dependent ArfGAP proteins 2 and 3 have preferences to binding individual coat complexes. Likewise, comparison of binding of the ArfGAPs to the soluble and the polymerized states of the coat protein will allow attribution of functions to the various ArfGAPs in uncoating of COPI vesicles.

What happens to coatomer during uncoating? Does hydrolysis of GTP by Arf1 leave a network of polymerized coatomer that falls off in pieces, and is the conformation of the complex reversed to its soluble form afterward by energy dependent chaperons, similar to the clathrin system (Schmid and Rothman 1985), or is each individual coatomer complex reversed in conformation during the uncoating reaction? Investigation of such questions may reveal novel proteins or roles of chaperons and will help finally establish the molecular mechanisms of the process of uncoating.

The small GTPase Arf1 takes part not only in the formation of the COPI coat, but is also a component of all types of clathrin-coated vesicles that bud from the endomembrane system but not the plasma membrane. What is the mechanism of membrane scission of these dynamin-independent vesicles? Does Arf1 play a role as a dimer in their scission reaction, similar to the COPI system?

We have learned about structures and functions of core components and individual basic molecular mechanisms of COPI budding in the Golgi, mainly from reconstituted systems in vitro. Although this is a prerequisite for our future integrated understanding, we need to learn a lot more about the roles of additional, only partly known, players as they operate in the context of a living cell.

Thus, investigating the biosynthesis of coated vesicular carriers will be a highly attractive challenge in the future for biochemists, structural biologists, and cell biologists.

ACKNOWLEDGMENTS

We apologize to those colleagues whose work we could not cite because of the limited space. We thank Thomas Soellner and Patricia McCabe for critical reading the manuscript, and the German Research Council for supporting our work (SFB 638, projects A10 and A16, and GRK 1188). V.P. is supported by a FEBS long-term fellowship.

Footnotes

Editors: Graham Warren and James Rothman

Additional Perspectives on The Golgi available at www.cshperspectives.org

REFERENCES

  • Aguilera-Romero A, Kaminska J, Spang A, Riezman H, Muniz M 2008. The yeast p24 complex is required for the formation of COPI retrograde transport vesicles from the Golgi apparatus. J Cell Biol 180: 713–720 [PMC free article] [PubMed] [Google Scholar]
  • Andag U, Schmitt HD 2003. Dsl1p, an essential component of the Golgi–endoplasmic reticulum retrieval system in yeast, uses the same sequence motif to interact with different subunits of the COPI vesicle coat. J Biol Chem 278: 51722–51734 [PubMed] [Google Scholar]
  • Andag U, Neumann T, Schmitt HD 2001. The coatomer-interacting protein Dsl1p is required for Golgi-to-endoplasmic reticulum retrieval in yeast. J Biol Chem 276: 39150–39160 [PubMed] [Google Scholar]
  • Aniento F, Gu F, Parton RG, Gruenberg J 1996. An endosomal beta COP is involved in the pH–dependent formation of transport vesicles destined for late endosomes. J Cell Biol 133: 29–41 [PMC free article] [PubMed] [Google Scholar]
  • Antonny B, Beraud-Dufour S, Chardin P, Chabre M 1997. N-terminal hydrophobic residues of the G-protein ADP-ribosylation factor-1 insert into membrane phospholipids upon GDP to GTP exchange. Biochemistry 36: 4675–4684 [PubMed] [Google Scholar]
  • Aoe T, Cukierman E, Lee A, Cassel D, Peters PJ, Hsu VW 1997. The KDEL receptor, ERD2, regulates intracellular traffic by recruiting a GTPase-activating protein for ARF1. EMBO J 16: 7305–7316 [PMC free article] [PubMed] [Google Scholar]
  • Aoe T, Lee AJ, van Donselaar E, Peters PJ, Hsu VW 1998. Modulation of intracellular transport by transported proteins: Insight from regulation of COPI-mediated transport. Proc Natl Acad Sci 95: 1624–1629 [PMC free article] [PubMed] [Google Scholar]
  • Aridor M, Bannykh SI, Rowe T, Balch WE 1995. Sequential coupling between COPII and COPI vesicle coats in endoplasmic reticulum to Golgi transport. J Cell Biol 131: 875–893 [PMC free article] [PubMed] [Google Scholar]
  • Asp L, Kartberg F, Fernandez-Rodriguez J, Smedh M, Elsner M, Laporte F, Barcena M, Jansen KA, Valentijn JA, Koster AJ, et al. 2009. Early stages of Golgi vesicle and tubule formation require diacylglycerol. Mol Biol Cell 20: 780–790 [PMC free article] [PubMed] [Google Scholar]
  • Barlowe C, Orci L, Yeung T, Hosobuchi M, Hamamoto S, Salama N, Rexach MF, Ravazzola M, Amherdt M, Schekman R 1994. COPII: A membrane coat formed by Sec proteins that drive vesicle budding from the endoplasmic reticulum. Cell 77: 895–907 [PubMed] [Google Scholar]
  • Bashkirov PV, Akimov SA, Evseev AI, Schmid SL, Zimmerberg J, Frolov VA 2008. GTPase cycle of dynamin is coupled to membrane squeeze and release, leading to spontaneous fission. Cell 135: 1276–1286 [PMC free article] [PubMed] [Google Scholar]
  • Beck R, Sun Z, Adolf F, Rutz C, Bassler J, Wild K, Sinning I, Hurt E, Brugger B, Bethune J, et al. 2008. Membrane curvature induced by Arf1-GTP is essential for vesicle formation. Proc Natl Acad Sci 105: 11731–11736 [PMC free article] [PubMed] [Google Scholar]
  • Beck R, Adolf F, Weimer C, Bruegger B, Wieland FT 2009a. ArfGAP1 activity and COPI vesicle biogenesis. Traffic 10: 307–315 [PubMed] [Google Scholar]
  • Beck R, Rawet M, Wieland FT, Cassel D 2009b. The COPI system: Molecular mechanisms and function. FEBS Lett 583: 2701–2709 [PubMed] [Google Scholar]
  • Beller M, Sztalryd C, Southall N, Bell M, Jackle H, Auld DS, Oliver B 2008. COPI complex is a regulator of lipid homeostasis. PLoS Biol 6: e292. [PMC free article] [PubMed] [Google Scholar]
  • Beraud-Dufour S, Robineau S, Chardin P, Paris S, Chabre M, Cherfils J, Antonny B 1998. A glutamic finger in the guanine nucleotide exchange factor ARNO displaces Mg2+ and the beta-phosphate to destabilize GDP on ARF1. EMBO J 17: 3651–3659 [PMC free article] [PubMed] [Google Scholar]
  • Bermak JC, Li M, Bullock C, Weingarten P, Zhou QY 2002. Interaction of gamma-COP with a transport motif in the D1 receptor C-terminus. Eur J Cell Biol 81: 77–85 [PubMed] [Google Scholar]
  • Beron W, Mayorga LS, Colombo MI, Stahl PD 2001. Recruitment of coat-protein-complex proteins on to phagosomal membranes is regulated by a brefeldin A-sensitive ADP-ribosylation factor. Biochem J 355: 409–415 [PMC free article] [PubMed] [Google Scholar]
  • Bethune J, Kol M, Hoffmann J, Reckmann I, Brugger B, Wieland F 2006. Coatomer, the coat protein of COPI transport vesicles, discriminates endoplasmic reticulum residents from p24 proteins. Mol Cell Biol 26: 8011–8021 [PMC free article] [PubMed] [Google Scholar]
  • Bielli A, Haney CJ, Gabreski G, Watkins SC, Bannykh SI, Aridor M 2005. Regulation of Sar1 NH2 terminus by GTP binding and hydrolysis promotes membrane deformation to control COPII vesicle fission. J Cell Biol 171: 919–924 [PMC free article] [PubMed] [Google Scholar]
  • Bigay J, Gounon P, Robineau S, Antonny B 2003. Lipid packing sensed by ArfGAP1 couples COPI coat disassembly to membrane bilayer curvature. Nature 426: 563–566 [PubMed] [Google Scholar]
  • Bigay J, Casella JF, Drin G, Mesmin B, Antonny B 2005. ArfGAP1 responds to membrane curvature through the folding of a lipid packing sensor motif. EMBO J 24: 2244–2253 [PMC free article] [PubMed] [Google Scholar]
  • Bonifacino JS, Hurley JH 2008. Retromer. Curr Opin Cell Biol 8: 8 [PMC free article] [PubMed] [Google Scholar]
  • Botelho RJ, Hackam DJ, Schreiber AD, Grinstein S 2000. Role of COPI in phagosome maturation. J Biol Chem 275: 15717–15727 [PubMed] [Google Scholar]
  • Bremser M, Nickel W, Schweikert M, Ravazzola M, Amherdt M, Hughes CA, Sollner TH, Rothman JE, Wieland FT 1999. Coupling of coat assembly and vesicle budding to packaging of putative cargo receptors. Cell 96: 495–506 [PubMed] [Google Scholar]
  • Brugger B, Sandhoff R, Wegehingel S, Gorgas K, Malsam J, Helms JB, Lehmann WD, Nickel W, Wieland FT 2000. Evidence for segregation of sphingomyelin and cholesterol during formation of COPI-coated vesicles. J Cell Biol 151: 507–518 [PMC free article] [PubMed] [Google Scholar]
  • Cabrera M, Muniz M, Hidalgo J, Vega L, Martin ME, Velasco A 2003. The retrieval function of the KDEL receptor requires PKA phosphorylation of its C-terminus. Mol Biol Cell 14: 4114–4125 [PMC free article] [PubMed] [Google Scholar]
  • Carvou N, Holic R, Li M, Futter C, Skippen A, Cockcroft S 2010. Phosphatidylinositol- and phosphatidylcholine-transfer activity of PITPbeta is essential for COPI-mediated retrograde transport from the Golgi to the endoplasmic reticulum. J Cell Sci 123: 1262–1273 [PMC free article] [PubMed] [Google Scholar]
  • Casanova JE 2007. Regulation of Arf activation: The Sec7 family of guanine nucleotide exchange factors. Traffic 8: 1476–1485 [PubMed] [Google Scholar]
  • Chen JL, Fucini RV, Lacomis L, Erdjument-Bromage H, Tempst P, Stamnes M 2005. Coatomer-bound Cdc42 regulates dynein recruitment to COPI vesicles. J Cell Biol 169: 383–389 [PMC free article] [PubMed] [Google Scholar]
  • Claude A, Zhao BP, Kuziemsky CE, Dahan S, Berger SJ, Yan JP, Armold AD, Sullivan EM, Melancon P 1999. GBF1: A novel Golgi-associated BFA-resistant guanine nucleotide exchange factor that displays specificity for ADP-ribosylation factor 5. J Cell Biol 146: 71–84 [PMC free article] [PubMed] [Google Scholar]
  • Cosson P, Letourneur F 1994. Coatomer interaction with di-lysine endoplasmic reticulum retention motifs. Science 263: 1629–1631 [PubMed] [Google Scholar]
  • Cosson P, Lefkir Y, Demolliere C, Letourneur F 1998. New COP1-binding motifs involved in ER retrieval. EMBO J 17: 6863–6870 [PMC free article] [PubMed] [Google Scholar]
  • Cukierman E, Huber I, Rotman M, Cassel D 1995. The ARF1 GTPase- activating protein: Zinc finger motif and Golgi complex localization. Science 270: 1999–2002 [PubMed] [Google Scholar]
  • Daro E, Sheff D, Gomez M, Kreis T, Mellman I 1997. Inhibition of endosome function in CHO cells bearing a temperature-sensitive defect in the coatomer (COPI) component epsilon-COP. J Cell Biol 139: 1747–1759 [PMC free article] [PubMed] [Google Scholar]
  • Dell’Angelica EC, Mullins C, Bonifacino JS 1999. AP-4, a novel protein complex related to clathrin adaptors. J Biol Chem 274: 7278–7285 [PubMed] [Google Scholar]
  • Denzel A, Otto F, Girod A, Pepperkok R, Watson R, Rosewell I, Bergeron JJ, Solari RC, Owen MJ 2000. The p24 family member p23 is required for early embryonic development. Curr Biol 10: 55–58 [PubMed] [Google Scholar]
  • Donaldson JG, Kahn RA, Lippincott-Schwartz J, Klausner RD 1991. Binding of ARF and beta-COP to Golgi membranes: Possible regulation by a trimeric G protein. Science 254: 1197–1199 [PubMed] [Google Scholar]
  • Donaldson JG, Honda A, Weigert R 2005. Multiple activities for Arf1 at the Golgi complex. Biochim Biophys Acta 1744: 364–373 [PubMed] [Google Scholar]
  • Duden R, Griffiths G, Frank R, Argos P, Kreis TE 1991. Beta-COP, a 110 kd protein associated with non-clathrin-coated vesicles and the Golgi complex, shows homology to beta-adaptin. Cell 64: 649–665 [PubMed] [Google Scholar]
  • Dumaresq-Doiron K, Savard MF, Akam S, Costantino S, Lefrancois S 2010. The phosphatidylinositol 4-kinase PI4KIIIalpha is required for the recruitment of GBF1 to Golgi membranes. J Cell Sci 123: 2273–2280 [PubMed] [Google Scholar]
  • East MP, Kahn RA 2010. Models for the functions of Arf GAPs. Semin Cell Dev Biol 22: 3–9 [PMC free article] [PubMed] [Google Scholar]
  • Eugster A, Frigerio G, Dale M, Duden R 2000. COP I domains required for coatomer integrity, and novel interactions with ARF and ARF-GAP. EMBO J 19: 3905–3917 [PMC free article] [PubMed] [Google Scholar]
  • Eugster A, Frigerio G, Dale M, Duden R 2004. The α- and β′-COP WD40 domains mediate cargo-selective interactions with distinct di-lysine motifs. Mol Biol Cell 15: 1011–1023 [PMC free article] [PubMed] [Google Scholar]
  • Farmaki T, Ponnambalam S, Prescott AR, Clausen H, Tang BL, Hong W, Lucocq JM 1999. Forward and retrograde trafficking in mitotic animal cells. ER-Golgi transport arrest restricts protein export from the ER into COPII-coated structures. J Cell Sci 112: 589–600 [PubMed] [Google Scholar]
  • Fernandez-Ulibarri I, Vilella M, Lazaro-Dieguez F, Sarri E, Martinez SE, Jimenez N, Claro E, Merida I, Burger KN, Egea G 2007. Diacylglycerol is required for the formation of COPI vesicles in the Golgi-to-ER transport pathway. Mol Biol Cell 18: 3250–3263 [PMC free article] [PubMed] [Google Scholar]
  • Fiedler K, Veit M, Stamnes MA, Rothman JE 1996a. Bimodal interaction of coatomer with the p24 family of putative cargo receptors. Science 273: 1396–1399 [PubMed] [Google Scholar]
  • Fiedler K, Veit M, Stamnes MA, Rothman JE 1996b. Bimodal interaction of coatomer with the p24 family of putative cargo receptors. Science 273: 1396–1399 [PubMed] [Google Scholar]
  • Franco M, Chardin P, Chabre M, Paris S 1996. Myristoylation-facilitated binding of the G protein ARF1GDP to membrane phospholipids is required for its activation by a soluble nucleotide exchange factor. J Biol Chem 271: 1573–1578 [PubMed] [Google Scholar]
  • Frigerio G, Grimsey N, Dale M, Majoul I, Duden R 2007. Two human ARFGAPs associated with COP-I-coated vesicles. Traffic 8: 1644–1655 [PMC free article] [PubMed] [Google Scholar]
  • Futatsumori M, Kasai K, Takatsu H, Shin HW, Nakayama K 2000. Identification and characterization of novel isoforms of COP I subunits. J Biochem 128: 793–801 [PubMed] [Google Scholar]
  • Gabriely G, Kama R, Gerst JE 2007. Involvement of specific COPI subunits in protein sorting from the late endosome to the vacuole in yeast. Mol Cell Biol 27: 526–540 [PMC free article] [PubMed] [Google Scholar]
  • Gallop JL, Butler PJ, McMahon HT 2005. Endophilin and CtBP/BARS are not acyl transferases in endocytosis or Golgi fission. Nature 438: 675–678 [PubMed] [Google Scholar]
  • Garcia-Mata R, Szul T, Alvarez C, Sztul E 2003. ADP-ribosylation factor/COPI-dependent events at the endoplasmic reticulum-Golgi interface are regulated by the guanine nucleotide exchange factor GBF1. Mol Biol Cell 14: 2250–2261 [PMC free article] [PubMed] [Google Scholar]
  • Goldberg J 1999. Structural and functional analysis of the ARF1-ARFGAP complex reveals a role for coatomer in GTP hydrolysis. Cell 96: 893–902 [PubMed] [Google Scholar]
  • Goldberg J 2000. Decoding of sorting signals by coatomer through a GTPase switch in the COPI coat complex. Cell 100: 671–679 [PubMed] [Google Scholar]
  • Gommel DU, Memon AR, Heiss A, Lottspeich F, Pfannstiel J, Lechner J, Reinhard C, Helms JB, Nickel W, Wieland FT 2001. Recruitment to Golgi membranes of ADP-ribosylation factor 1 is mediated by the cytoplasmic domain of p23. EMBO J 20: 6751–6760 [PMC free article] [PubMed] [Google Scholar]
  • Griffiths G, Pepperkok R, Locker JK, Kreis TE 1995. Immunocytochemical localization of beta-COP to the ER-Golgi boundary and the TGN. J Cell Sci 108: 2839–2856 [PubMed] [Google Scholar]
  • Gu F, Gruenberg J 2000. ARF1 regulates pH-dependent COP functions in the early endocytic pathway. J Biol Chem 275: 8154–8160 [PubMed] [Google Scholar]
  • Gu F, Aniento F, Parton RG, Gruenberg J 1997. Functional dissection of COP-I subunits in the biogenesis of multivesicular endosomes. J Cell Biol 139: 1183–1195 [PMC free article] [PubMed] [Google Scholar]
  • Guo Y, Walther TC, Rao M, Stuurman N, Goshima G, Terayama K, Wong JS, Vale RD, Walter P, Farese RV 2008. Functional genomic screen reveals genes involved in lipid-droplet formation and utilization. Nature 453: 657–661 [PMC free article] [PubMed] [Google Scholar]
  • Hackam DJ, Botelho RJ, Sjolin C, Rotstein OD, Robinson JM, Schreiber AD, Grinstein S 2001. Indirect role for COPI in the completion of FCgamma receptor-mediated phagocytosis. J Biol Chem 276: 18200–18208 [PubMed] [Google Scholar]
  • Hara-Kuge S, Kuge O, Orci L, Amherdt M, Ravazzola M, Wieland FT, Rothman JE 1994. En bloc incorporation of coatomer subunits during the assembly of COP-coated vesicles. J Cell Biol 124: 883–892 [PMC free article] [PubMed] [Google Scholar]
  • Harrison-Lavoie KJ, Lewis VA, Hynes GM, Collison KS, Nutland E, Willison KR 1993. A 102 kDa subunit of a Golgi-associated particle has homology to beta subunits of trimeric G proteins. EMBO J 12: 2847–2853 [PMC free article] [PubMed] [Google Scholar]
  • Harter C, Wieland FT 1998. A single binding site for dilysine retrieval motifs and p23 within the gamma subunit of coatomer. Proc Natl Acad Sci 95: 11649–11654 [PMC free article] [PubMed] [Google Scholar]
  • Hehnly H, Xu W, Chen JL, Stamnes M 2010. Cdc42 regulates microtubule-dependent Golgi positioning. Traffic 11: 1067–1078 [PMC free article] [PubMed] [Google Scholar]
  • Hirst J, Lui WW, Bright NA, Totty N, Seaman MN, Robinson MS 2000. A family of proteins with gamma-adaptin and VHS domains that facilitate trafficking between the trans-Golgi network and the vacuole/lysosome. J Cell Biol 149: 67–80 [PMC free article] [PubMed] [Google Scholar]
  • Hoffman GR, Rahl PB, Collins RN, Cerione RA 2003. Conserved structural motifs in intracellular trafficking pathways: structure of the gammaCOP appendage domain. Mol Cell 12: 615–625 [PubMed] [Google Scholar]
  • Honda A, Al-Awar OS, Hay JC, Donaldson JG 2005. Targeting of Arf-1 to the early Golgi by membrin, an ER-Golgi SNARE. J Cell Biol 168: 1039–1051 [PMC free article] [PubMed] [Google Scholar]
  • Hsia KC, Hoelz A 2010. Crystal structure of alpha-COP in complex with epsilon-COP provides insight into the architecture of the COPI vesicular coat. Proc Natl Acad Sci 107: 11271–11276 [PMC free article] [PubMed] [Google Scholar]
  • Hsu VW, Yang JS 2009. Mechanisms of COPI vesicle formation. FEBS Lett 583: 3758–3763 [PMC free article] [PubMed] [Google Scholar]
  • Hughes H, Stephens DJ 2008. Assembly, organization, and function of the COPII coat. Histochem Cell Biol 129: 129–151 [PMC free article] [PubMed] [Google Scholar]
  • Hurley JH, Hanson PI 2010. Membrane budding and scission by the ESCRT machinery: It’s all in the neck. Nat Rev Mol Cell Biol 11: 556–566 [PMC free article] [PubMed] [Google Scholar]
  • Kartberg F, Asp L, Dejgaard SY, Smedh M, Fernandez-Rodriguez J, Nilsson T, Presley JF 2010. ARFGAP2 and ARFGAP3 are essential for copi coat assembly on the golgi membrane of living cells. J Biol Chem 285: 36709–36720 [PMC free article] [PubMed] [Google Scholar]
  • Katzmann DJ, Babst M, Emr SD 2001. Ubiquitin-dependent sorting into the multivesicular body pathway requires the function of a conserved endosomal protein sorting complex, ESCRT-I. Cell 106: 145–155 [PubMed] [Google Scholar]
  • Kawamoto K, Yoshida Y, Tamaki H, Torii S, Shinotsuka C, Yamashina S, Nakayama K 2002. GBF1, a guanine nucleotide exchange factor for ADP- ribosylation factors, is localized to the cis-Golgi and involved in membrane association of the COPI coat. Traffic 3: 483–495 [PubMed] [Google Scholar]
  • Klemm RW, Ejsing CS, Surma MA, Kaiser HJ, Gerl MJ, Sampaio JL, de Robillard Q, Ferguson C, Proszynski TJ, Shevchenko A, et al. 2009. Segregation of sphingolipids and sterols during formation of secretory vesicles at the trans-Golgi network. J Cell Biol 185: 601–612 [PMC free article] [PubMed] [Google Scholar]
  • Kliouchnikov L, Bigay J, Mesmin B, Parnis A, Rawet M, Goldfeder N, Antonny B, Cassel D 2009. Discrete determinants in ArfGAP2/3 conferring Golgi localization and regulation by the COPI coat. Mol Biol Cell 20: 859–869 [PMC free article] [PubMed] [Google Scholar]
  • Krauss M, Jia JY, Roux A, Beck R, Wieland FT, De Camilli P, Haucke V 2008. Arf1-GTP-induced tubule formation suggests a function of Arf family proteins in curvature acquisition at sites of vesicle budding. J Biol Chem 283: 27717–27723 [PMC free article] [PubMed] [Google Scholar]
  • Langer JD, Roth CM, Bethune J, Stoops EH, Brugger B, Herten DP, Wieland FT 2008. A conformational change in the alpha-subunit of coatomer induced by ligand binding to gamma-COP revealed by single- pair FRET. Traffic 9: 597–607 [PubMed] [Google Scholar]
  • Lanoix J, Ouwendijk J, Lin CC, Stark A, Love HD, Ostermann J, Nilsson T 1999. GTP hydrolysis by arf-1 mediates sorting and concentration of Golgi resident enzymes into functional COP I vesicles. EMBO J 18: 4935–4948 [PMC free article] [PubMed] [Google Scholar]
  • Lanoix J, Ouwendijk J, Stark A, Szafer E, Cassel D, Dejgaard K, Weiss M, Nilsson T 2001. Sorting of Golgi resident proteins into different subpopulations of COPI vesicles: A role for ArfGAP1. J Cell Biol 155: 1199–1212 [PMC free article] [PubMed] [Google Scholar]
  • Lay D, Gorgas K, Just WW 2006. Peroxisome biogenesis: Where Arf and coatomer might be involved. Biochim Biophys Acta 1763: 1678–1687 [PubMed] [Google Scholar]
  • Lee C, Goldberg J 2010. Structure of coatomer cage proteins and the relationship among COPI, COPII, and clathrin vesicle coats. Cell 142: 123–132 [PMC free article] [PubMed] [Google Scholar]
  • Lee MC, Miller EA, Goldberg J, Orci L, Schekman R 2004. Bi-directional protein transport between the ER and Golgi. Annu Rev Cell Dev Biol 20: 87–123 [PubMed] [Google Scholar]
  • Lee MC, Orci L, Hamamoto S, Futai E, Ravazzola M, Schekman R 2005. Sar1p N-terminal helix initiates membrane curvature and completes the fission of a COPII vesicle. Cell 122: 605–617 [PubMed] [Google Scholar]
  • Letourneur F, Gaynor EC, Hennecke S, Demolliere C, Duden R, Emr SD, Riezman H, Cosson P 1994. Coatomer is essential for retrieval of dilysine-tagged proteins to the endoplasmic reticulum. Cell 79: 1199–1207 [PubMed] [Google Scholar]
  • Lewis MJ, Pelham HR 1992. Ligand-induced redistribution of a human KDEL receptor from the Golgi complex to the endoplasmic reticulum. Cell 68: 353–364 [PubMed] [Google Scholar]
  • Liu J, Prunuske AJ, Fager AM, Ullman KS 2003. The COPI complex functions in nuclear envelope breakdown and is recruited by the nucleoporin Nup153. Dev Cell 5: 487–498 [PubMed] [Google Scholar]
  • Liu W, Duden R, Phair RD, Lippincott-Schwartz J 2005. ArfGAP1 dynamics and its role in COPI coat assembly on Golgi membranes of living cells. J Cell Biol 168: 1053–1063 [PMC free article] [PubMed] [Google Scholar]
  • Liu Y, Kahn RA, Prestegard JH 2010. Dynamic structure of membrane-anchored Arf*GTP. Nat Struct Mol Biol 17: 876–881 [PMC free article] [PubMed] [Google Scholar]
  • Long KR, Yamamoto Y, Baker AL, Watkins SC, Coyne CB, Conway JF, Aridor M 2010. Sar1 assembly regulates membrane constriction and ER export. J Cell Biol 190: 115–128 [PMC free article] [PubMed] [Google Scholar]
  • Lowe M, Kreis TE 1995. In vitro assembly and disassembly of coatomer. J Biol Chem 270: 31364–31371 [PubMed] [Google Scholar]
  • Lundmark R, Doherty GJ, Vallis Y, Peter BJ, McMahon HT 2008. Arf family GTP loading is activated by, and generates, positive membrane curvature. Biochem J 414: 189–194 [PMC free article] [PubMed] [Google Scholar]
  • Majoul I, Sohn K, Wieland FT, Pepperkok R, Pizza M, Hillemann J, Soling HD 1998. KDEL receptor (Erd2p)-mediated retrograde transport of the cholera toxin A subunit from the Golgi involves COPI, p23, and the COOH terminus of Erd2p. J Cell Biol 143: 601–612 [PMC free article] [PubMed] [Google Scholar]
  • Majoul I, Straub M, Hell SW, Duden R, Soling HD 2001. KDEL-cargo regulates interactions between proteins involved in COPI vesicle traffic: measurements in living cells using FRET. Dev Cell 1: 139–153 [PubMed] [Google Scholar]
  • Malhotra V, Serafini T, Orci L, Shepherd JC, Rothman JE 1989. Purification of a novel class of coated vesicles mediating biosynthetic protein transport through the Golgi stack. Cell 58: 329–336 [PubMed] [Google Scholar]
  • Malsam J, Sollner T 2011. Organization of SNAREs within the Golgi stack. Cold Spring Harb Perspect Biol 10.1101/cshperspect.a005249 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Malsam J, Gommel D, Wieland FT, Nickel W 1999. A role for ADP ribosylation factor in the control of cargo uptake during COPI-coated vesicle biogenesis. FEBS Lett 462: 267–272 [PubMed] [Google Scholar]
  • Malsam J, Satoh A, Pelletier L, Warren G 2005. Golgin tethers define subpopulations of COPI vesicles. Science 307: 1095–1098 [PubMed] [Google Scholar]
  • Manneville JB, Casella JF, Ambroggio E, Gounon P, Bertherat J, Bassereau P, Cartaud J, Antonny B, Goud B 2008. COPI coat assembly occurs on liquid-disordered domains and the associated membrane deformations are limited by membrane tension. Proc Natl Acad Sci 105: 16946–16951 [PMC free article] [PubMed] [Google Scholar]
  • Mansour SJ, Skaug J, Zhao XH, Giordano J, Scherer SW, Melancon P 1999. p200 ARF-GEP1: A Golgi-localized guanine nucleotide exchange protein whose Sec7 domain is targeted by the drug brefeldin A. Proc Natl Acad Sci 96: 7968–7973 [PMC free article] [PubMed] [Google Scholar]
  • McMahon HT, Mills IG 2004. COP and clathrin-coated vesicle budding: Different pathways, common approaches. Curr Opin Cell Biol 16: 379–391 [PubMed] [Google Scholar]
  • Michelsen K, Yuan H, Schwappach B 2005. Hide and run. Arginine-based endoplasmic-reticulum-sorting motifs in the assembly of heteromultimeric membrane proteins. EMBO Rep 6: 717–722 [PMC free article] [PubMed] [Google Scholar]
  • Michelsen K, Schmid V, Metz J, Heusser K, Liebel U, Schwede T, Spang A, Schwappach B 2007. Novel cargo-binding site in the beta and delta subunits of coatomer. J Cell Biol 179: 209–217 [PMC free article] [PubMed] [Google Scholar]
  • Misteli T, Warren G 1994. COP-coated vesicles are involved in the mitotic fragmentation of Golgi stacks in a cell-free system. J Cell Biol 125: 269–282 [PMC free article] [PubMed] [Google Scholar]
  • Moelleken J, Malsam J, Betts MJ, Movafeghi A, Reckmann I, Meissner I, Hellwig A, Russell RB, Sollner T, Brugger B, et al. 2007. Differential localization of coatomer complex isoforms within the Golgi apparatus. Proc Natl Acad Sci 104: 4425–4430 [PubMed] [Google Scholar]
  • Monetta P, Slavin I, Romero N, Alvarez C 2007. Rab1b interacts with GBF1 and modulates both ARF1 dynamics and COPI association. Mol Biol Cell 18: 2400–2410 [PMC free article] [PubMed] [Google Scholar]
  • Murphy JE, Pleasure IT, Puszkin S, Prasad K, Keen JH 1991. Clathrin assembly protein AP-3. The identity of the 155K protein, AP 180, and NP185 and demonstration of a clathrin binding domain. J Biol Chem 266: 4401–4408 [PubMed] [Google Scholar]
  • Nachury MV, Loktev AV, Zhang Q, Westlake CJ, Peranen J, Merdes A, Slusarski DC, Scheller RH, Bazan JF, Sheffield VC, et al. 2007. A core complex of BBS proteins cooperates with the GTPase Rab8 to promote ciliary membrane biogenesis. Cell 129: 1201–1213 [PubMed] [Google Scholar]
  • Nachury MV, Seeley ES, Jin H 2010. Trafficking to the ciliary membrane: How to get across the periciliary diffusion barrier? Annu Rev Cell Dev Biol 26: 59–87 [PMC free article] [PubMed] [Google Scholar]
  • Nickel W, Malsam J, Gorgas K, Ravazzola M, Jenne N, Helms JB, Wieland FT 1998. Uptake by COPI-coated vesicles of both anterograde and retrograde cargo is inhibited by GTPgammaS in vitro. J Cell Sci 111: 3081–3090 [PubMed] [Google Scholar]
  • Nie Z, Hirsch DS, Randazzo PA 2003. Arf and its many interactors. Curr Opin Cell Biol 15: 396–404 [PubMed] [Google Scholar]
  • Nilsson T, Jackson M, Peterson PA 1989. Short cytoplasmic sequences serve as retention signals for transmembrane proteins in the endoplasmic reticulum. Cell 58: 707–718 [PubMed] [Google Scholar]
  • Oka T, Nakano A 1994. Inhibition of GTP hydrolysis by Sar1p causes accumulation of vesicles that are a functional intermediate of the ER-to- Golgi transport in yeast. J Cell Biol 124: 425–434 [PMC free article] [PubMed] [Google Scholar]
  • O’Kelly I, Butler MH, Zilberberg N, Goldstein SA 2002. Forward transport. 14-3-3 binding overcomes retention in endoplasmic reticulum by dibasic signals. Cell 111: 577–588 [PubMed] [Google Scholar]
  • Orci L, Glick BS, Rothman JE 1986. A new type of coated vesicular carrier that appears not to contain clathrin: Its possible role in protein transport within the Golgi stack. Cell 46: 171–184 [PubMed] [Google Scholar]
  • Orci L, Palmer DJ, Ravazzola M, Perrelet A, Amherdt M, Rothman JE 1993. Budding from Golgi membranes requires the coatomer complex of non-clathrin coat proteins. Nature 362: 648–652 [PubMed] [Google Scholar]
  • Orci L, Perrelet A, Ravazzola M, Amherdt M, Rothman JE, Schekman R 1994. Coatomer-rich endoplasmic reticulum. Proc Natl Acad Sci 91: 11924–11928 [PMC free article] [PubMed] [Google Scholar]
  • Orci L, Stamnes M, Ravazzola M, Amherdt M, Perrelet A, Sollner TH, Rothman JE 1997. Bidirectional transport by distinct populations of COPI-coated vesicles. Cell 90: 335–349 [PubMed] [Google Scholar]
  • Orci L, Ravazzola M, Volchuk A, Engel T, Gmachl M, Amherdt M, Perrelet A, Sollner TH, Rothman JE 2000. Anterograde flow of cargo across the golgi stack potentially mediated via bidirectional “percolating” COPI vesicles. Proc Natl Acad Sci 97: 10400–10405 [PMC free article] [PubMed] [Google Scholar]
  • Palmer DJ, Helms JB, Beckers CJ, Orci L, Rothman JE 1993. Binding of coatomer to Golgi membranes requires ADP-ribosylation factor. J Biol Chem 268: 12083–12089 [PubMed] [Google Scholar]
  • Pavel J, Harter C, Wieland FT 1998. Reversible dissociation of coatomer: Functional characterization of a beta/delta-coat protein subcomplex. Proc Natl Acad Sci 95: 2140–2145 [PMC free article] [PubMed] [Google Scholar]
  • Pearse BM 1976. Clathrin: A unique protein associated with intracellular transfer of membrane by coated vesicles. Proc Natl Acad Sci 73: 1255–1259 [PMC free article] [PubMed] [Google Scholar]
  • Pearse BM, Robinson MS 1984. Purification and properties of 100-kd proteins from coated vesicles and their reconstitution with clathrin. EMBO J 3: 1951–1957 [PMC free article] [PubMed] [Google Scholar]
  • Pelham HR 1991. Recycling of proteins between the endoplasmic reticulum and Golgi complex. Curr Opin Cell Biol 3: 585–591 [PubMed] [Google Scholar]
  • Pelham HR 1994. About turn for the COPs? Cell 79: 1125–1127 [PubMed] [Google Scholar]
  • Pepperkok R, Whitney JA, Gomez M, Kreis TE 2000. COPI vesicles accumulating in the presence of a GTP restricted arf1 mutant are depleted of anterograde and retrograde cargo. J Cell Sci 113: 135–144 [PubMed] [Google Scholar]
  • Pinot M, Goud B, Manneville JB 2010. Physical aspects of COPI vesicle formation. Mol Membr Biol 27: 428–442 [PubMed] [Google Scholar]
  • Poon PP, Cassel D, Spang A, Rotman M, Pick E, Singer RA, Johnston GC 1999. Retrograde transport from the yeast Golgi is mediated by two ARF GAP proteins with overlapping function. EMBO J 18: 555–564 [PMC free article] [PubMed] [Google Scholar]
  • Presley JF, Ward TH, Pfeifer AC, Siggia ED, Phair RD, Lippincott-Schwartz J 2002. Dissection of COPI and Arf1 dynamics in vivo and role in Golgi membrane transport. Nature 417: 187–193 [PubMed] [Google Scholar]
  • Pucadyil TJ, Schmid SL 2008. Real-time visualization of dynamin-catalyzed membrane fission and vesicle release. Cell 135: 1263–1275 [PMC free article] [PubMed] [Google Scholar]
  • Randazzo PA, Yang YC, Rulka C, Kahn RA 1993. Activation of ADP- ribosylation factor by Golgi membranes. Evidence for a brefeldin A- and protease-sensitive activating factor on Golgi membranes. J Biol Chem 268: 9555–9563 [PubMed] [Google Scholar]
  • Razi M, Chan EY, Tooze SA 2009. Early endosomes and endosomal coatomer are required for autophagy. J Cell Biol 185: 305–321 [PMC free article] [PubMed] [Google Scholar]
  • Reilly BA, Kraynack BA, VanRheenen SM, Waters MG 2001. Golgi-to-endoplasmic reticulum (ER) retrograde traffic in yeast requires Dsl1p, a component of the ER target site that interacts with a COPI coat subunit. Mol Biol Cell 12: 3783–3796 [PMC free article] [PubMed] [Google Scholar]
  • Reinhard C, Harter C, Bremser M, Brugger B, Sohn K, Helms JB, Wieland F 1999. Receptor-induced polymerization of coatomer. Proc Natl Acad Sci 96: 1224–1228 [PMC free article] [PubMed] [Google Scholar]
  • Reinhard C, Schweikert M, Wieland FT, Nickel W 2003. Functional reconstitution of COPI coat assembly and disassembly using chemically defined components. Proc Natl Acad Sci 100: 8253–8257 [PMC free article] [PubMed] [Google Scholar]
  • Robinson M 2004. Adaptable adaptors for coated vesicles. Trends Cell Biol 14: 167–174 [PubMed] [Google Scholar]
  • Saitoh A, Shin HW, Yamada A, Waguri S, Nakayama K 2009. Three homologous ArfGAPs participate in coat protein I-mediated transport. J Biol Chem 284: 13948–13957 [PMC free article] [PubMed] [Google Scholar]
  • Sato K, Nakano A 2005. Dissection of COPII subunit-cargo assembly and disassembly kinetics during Sar1p-GTP hydrolysis. Nat Struct Mol Biol 12: 167–174 [PubMed] [Google Scholar]
  • Sato K, Nakano A 2007. Mechanisms of COPII vesicle formation and protein sorting. FEBS Lett 581: 2076–2082 [PubMed] [Google Scholar]
  • Scales SJ, Pepperkok R, Kreis TE 1997. Visualization of ER-to-Golgi transport in living cells reveals a sequential mode of action for COPII and COPI. Cell 90: 1137–1148 [PubMed] [Google Scholar]
  • Schmid SL, Rothman JE 1985. Two classes of binding sites for uncoating protein in clathrin triskelions. J Biol Chem 260: 10050–10056 [PubMed] [Google Scholar]
  • Schmitz KR, Liu J, Li S, Setty TG, Wood CS, Burd CG, Ferguson KM 2008. Golgi localization of glycosyltransferases requires a Vps74p oligomer. Dev Cell 14: 523–534 [PMC free article] [PubMed] [Google Scholar]
  • Schroder-Kohne S, Letourneur F, Riezman H 1998. Alpha-COP can discriminate between distinct, functional di-lysine signals in vitro and regulates access into retrograde transport. J Cell Sci 111: 3459–3470 [PubMed] [Google Scholar]
  • Scott DB, Blanpied TA, Swanson GT, Zhang C, Ehlers MD 2001. An NMDA receptor ER retention signal regulated by phosphorylation and alternative splicing. J Neurosci 21: 3063–3072 [PMC free article] [PubMed] [Google Scholar]
  • Seaman MN, McCaffery JM, Emr SD 1998. A membrane coat complex essential for endosome-to-Golgi retrograde transport in yeast. J Cell Biol 142: 665–681 [PMC free article] [PubMed] [Google Scholar]
  • Serafini T, Orci L, Amherdt M, Brunner M, Kahn RA, Rothman JE 1991a. ADP-ribosylation factor is a subunit of the coat of Golgi-derived COP-coated vesicles: A novel role for a GTP-binding protein. Cell 67: 239–253 [PubMed] [Google Scholar]
  • Serafini T, Stenbeck G, Brecht A, Lottspeich F, Orci L, Rothman JE, Wieland FT 1991b. A coat subunit of Golgi-derived non-clathrin-coated vesicles with homology to the clathrin-coated vesicle coat protein beta- adaptin. Nature 349: 215–220 [PubMed] [Google Scholar]
  • Shikano S, Li M 2003. Membrane receptor trafficking: Evidence of proximal and distal zones conferred by two independent endoplasmic reticulum localization signals. Proc Natl Acad Sci 100: 5783–5788 [PMC free article] [PubMed] [Google Scholar]
  • Shima DT, Scales SJ, Kreis TE, Pepperkok R 1999. Segregation of COPI- rich and anterograde-cargo-rich domains in endoplasmic-reticulum-to-Golgi transport complexes. Curr Biol 9: 821–824 [PubMed] [Google Scholar]
  • Shin HW, Morinaga N, Noda M, Nakayama K 2004. BIG2, a guanine nucleotide exchange factor for ADP-ribosylation factors: Its localization to recycling endosomes and implication in the endosome integrity. Mol Biol Cell 15: 5283–5294 [PMC free article] [PubMed] [Google Scholar]
  • Shinotsuka C, Yoshida Y, Kawamoto K, Takatsu H, Nakayama K 2002. Overexpression of an ADP-ribosylation factor-guanine nucleotide exchange factor, BIG2, uncouples brefeldin A-induced adaptor protein-1 coat dissociation and membrane tubulation. J Biol Chem 277: 9468–9473 [PubMed] [Google Scholar]
  • Simon JP, Morimoto T, Bankaitis VA, Gottlieb TA, Ivanov IE, Adesnik M, Sabatini DD 1998. An essential role for the phosphatidylinositol transfer protein in the scission of coatomer-coated vesicles from the trans-Golgi network. Proc Natl Acad Sci 95: 11181–11186 [PMC free article] [PubMed] [Google Scholar]
  • Sohn K, Orci L, Ravazzola M, Amherdt M, Bremser M, Lottspeich F, Fiedler K, Helms JB, Wieland FT 1996. A major transmembrane protein of Golgi-derived COPI-coated vesicles involved in coatomer binding. J Cell Biol 135: 1239–1248 [PMC free article] [PubMed] [Google Scholar]
  • Soni KG, Mardones GA, Sougrat R, Smirnova E, Jackson CL, Bonifacino JS 2009. Coatomer-dependent protein delivery to lipid droplets. J Cell Sci 122: 1834–1841 [PMC free article] [PubMed] [Google Scholar]
  • Spang A, Matsuoka K, Hamamoto S, Schekman R, Orci L 1998. Coatomer, Arf1p, and nucleotide are required to bud coat protein complex I-coated vesicles from large synthetic liposomes. Proc Natl Acad Sci 95: 11199–11204 [PMC free article] [PubMed] [Google Scholar]
  • Spang A, Herrmann JM, Hamamoto S, Schekman R 2001. The ADP ribosylation factor-nucleotide exchange factors Gea1p and Gea2p have overlapping, but not redundant functions in retrograde transport from the Golgi to the endoplasmic reticulum. Mol Biol Cell 12: 1035–1045 [PMC free article] [PubMed] [Google Scholar]
  • Springer S, Chen E, Duden R, Marzioch M, Rowley A, Hamamoto S, Merchant S, Schekman R 2000. The p24 proteins are not essential for vesicular transport in Saccharomyces cerevisiae. Proc Natl Acad Sci 97: 4034–4039 [PMC free article] [PubMed] [Google Scholar]
  • Stamnes MA, Craighead MW, Hoe MH, Lampen N, Geromanos S, Tempst P, Rothman JE 1995. An integral membrane component of coatomer- coated transport vesicles defines a family of proteins involved in budding. Proc Natl Acad Sci 92: 8011–8015 [PMC free article] [PubMed] [Google Scholar]
  • Standley S, Roche KW, McCallum J, Sans N, Wenthold RJ 2000. PDZ domain suppression of an ER retention signal in NMDA receptor NR1 splice variants. Neuron 28: 887–898 [PubMed] [Google Scholar]
  • Stenbeck G, Harter C, Brecht A, Herrmann D, Lottspeich F, Orci L, Wieland FT 1993. beta'-COP, a novel subunit of coatomer. EMBO J 12: 2841–2845 [PMC free article] [PubMed] [Google Scholar]
  • Sun Z, Anderl F, Frohlich K, Zhao L, Hanke S, Brugger B, Wieland F, Bethune J 2007. Multiple and stepwise interactions between coatomer and ADP-ribosylation factor-1 (Arf1)-GTP. Traffic 8: 582–593 [PubMed] [Google Scholar]
  • Szafer E, Rotman M, Cassel D 2001. Regulation of GTP hydrolysis on ADP-ribosylation factor-1 at the Golgi membrane. J Biol Chem 276: 47834–47839 [PubMed] [Google Scholar]
  • Tabata KV, Sato K, Ide T, Nishizaka T, Nakano A, Noji H 2009. Visualization of cargo concentration by COPII minimal machinery in a planar lipid membrane. EMBO J 28: 3279–3289 [PMC free article] [PubMed] [Google Scholar]
  • Takamori S, Holt M, Stenius K, Lemke EA, Gronborg M, Riedel D, Urlaub H, Schenck S, Brugger B, Ringler P, et al. 2006. Molecular anatomy of a trafficking organelle. Cell 127: 831–846 [PubMed] [Google Scholar]
  • Takatsu H, Yoshino K, Nakayama K 2000. Adaptor gamma ear homology domain conserved in gamma-adaptin and GGA proteins that interact with gamma-synergin. Biochem Biophys Res Commun 271: 719–725 [PubMed] [Google Scholar]
  • Takeuchi M, Ueda T, Yahara N, Nakano A 2002. Arf1 GTPase plays roles in the protein traffic between the endoplasmic reticulum and the Golgi apparatus in tobacco and Arabidopsis cultured cells. Plant J 31: 499–515 [PubMed] [Google Scholar]
  • Tang D, Mar K, Warren G, Wang Y 2008. Molecular mechanism of mitotic Golgi disassembly and reassembly revealed by a defined reconstitution assay. J Biol Chem 283: 6085–6094 [PMC free article] [PubMed] [Google Scholar]
  • Tanigawa G, Orci L, Amherdt M, Ravazzola M, Helms JB, Rothman JE 1993. Hydrolysis of bound GTP by ARF protein triggers uncoating of Golgi-derived COP-coated vesicles. J Cell Biol 123: 1365–1371 [PMC free article] [PubMed] [Google Scholar]
  • Tripathi A, Ren Y, Jeffrey PD, Hughson FM 2009. Structural characterization of Tip20p and Dsl1p, subunits of the Dsl1p vesicle tethering complex. Nat Struct Mol Biol 16: 114–123 [PMC free article] [PubMed] [Google Scholar]
  • Tu L, Tai WC, Chen L, Banfield DK 2008. Signal-mediated dynamic retention of glycosyltransferases in the Golgi. Science 321: 404–407 [PubMed] [Google Scholar]
  • Volpicelli-Daley LA, Li Y, Zhang CJ, Kahn RA 2005. Isoform-selective effects of the depletion of ADP-ribosylation factors 1-5 on membrane traffic. Mol Biol Cell 16: 4495–4508 [PMC free article] [PubMed] [Google Scholar]
  • Wang CW, Hamamoto S, Orci L, Schekman R 2006. Exomer: A coat complex for transport of select membrane proteins from the trans-Golgi network to the plasma membrane in yeast. J Cell Biol 174: 973–983 [PMC free article] [PubMed] [Google Scholar]
  • Waters MG, Serafini T, Rothman JE 1991. “Coatomer”: A cytosolic protein complex containing subunits of non-clathrin-coated Golgi transport vesicles. Nature 349: 248–251 [PubMed] [Google Scholar]
  • Watson PJ, Frigerio G, Collins BM, Duden R, Owen DJ 2004. Gamma-COP appendage domain—structure and function. Traffic 5: 79–88 [PubMed] [Google Scholar]
  • Wegmann D, Hess P, Baier C, Wieland FT, Reinhard C 2004. Novel isotypic gamma/zeta subunits reveal three coatomer complexes in mammals. Mol Biol Cell 24: 1070–1080 [PMC free article] [PubMed] [Google Scholar]
  • Weimer C, Beck R, Eckert P, Reckmann I, Moelleken J, Brugger B, Wieland F 2008. Differential roles of ArfGAP1, ArfGAP2, and ArfGAP3 in COPI trafficking. J Cell Biol 183: 725–735 [PMC free article] [PubMed] [Google Scholar]
  • Weiss M, Nilsson T 2003. A kinetic proof-reading mechanism for protein sorting. Traffic 4: 65–73 [PubMed] [Google Scholar]
  • Whitney JA, Gomez M, Sheff D, Kreis TE, Mellman I 1995. Cytoplasmic coat proteins involved in endosome function. Cell 83: 703–713 [PubMed] [Google Scholar]
  • Wilson DW, Lewis MJ, Pelham HR 1993. pH-dependent binding of KDEL to its receptor in vitro. J Biol Chem 268: 7465–7468 [PubMed] [Google Scholar]
  • Wu WJ, Erickson JW, Lin R, Cerione RA 2000. The gamma-subunit of the coatomer complex binds Cdc42 to mediate transformation. Nature 405: 800–804 [PubMed] [Google Scholar]
  • Xing Y, Bocking T, Wolf M, Grigorieff N, Kirchhausen T, Harrison SC 2010. Structure of clathrin coat with bound Hsc70 and auxilin: Mechanism of Hsc70-facilitated disassembly. EMBO J 29: 655–665 [PMC free article] [PubMed] [Google Scholar]
  • Yang JS, Lee SY, Gao M, Bourgoin S, Randazzo PA, Premont RT, Hsu VW 2002. ARFGAP1 promotes the formation of COPI vesicles, suggesting function as a component of the coat. J Cell Biol 159: 69–78 [PMC free article] [PubMed] [Google Scholar]
  • Yang JS, Lee SY, Spano S, Gad H, Zhang L, Nie Z, Bonazzi M, Corda D, Luini A, Hsu VW 2005. A role for BARS at the fission step of COPI vesicle formation from Golgi membrane. EMBO J 24: 4133–4143 [PMC free article] [PubMed] [Google Scholar]
  • Yang JS, Zhang L, Lee SY, Gad H, Luini A, Hsu VW 2006. Key components of the fission machinery are interchangeable. Nat Cell Biol 8: 1376–1382 [PubMed] [Google Scholar]
  • Yang JS, Gad H, Lee SY, Mironov A, Zhang L, Beznoussenko GV, Valente C, Turacchio G, Bonsra AN, Du G, et al. 2008. A role for phosphatidic acid in COPI vesicle fission yields insights into Golgi maintenance. Nat Cell Biol 10: 1146–1153 [PMC free article] [PubMed] [Google Scholar]
  • Yu W, Lin J, Jin C, Xia B 2009. Solution structure of human zeta-COP: Direct evidences for structural similarity between COP I and clathrin- adaptor coats. J Mol Biol 386: 903–912 [PubMed] [Google Scholar]
  • Yuan H, Michelsen K, Schwappach B 2003. 14-3-3 dimers probe the assembly status of multimeric membrane proteins. Curr Biol 13: 638–646 [PubMed] [Google Scholar]
  • Zerangue N, Schwappach B, Jan YN, Jan LY 1999. A new ER trafficking signal regulates the subunit stoichiometry of plasma membrane K(ATP) channels. Neuron 22: 537–548 [PubMed] [Google Scholar]
  • Zerangue N, Malan MJ, Fried SR, Dazin PF, Jan YN, Jan LY, Schwappach B 2001. Analysis of endoplasmic reticulum trafficking signals by combinatorial screening in mammalian cells. Proc Natl Acad Sci 98: 2431–2436 [PMC free article] [PubMed] [Google Scholar]
  • Zhao L, Helms JB, Brugger B, Harter C, Martoglio B, Graf R, Brunner J, Wieland FT 1997. Direct and GTP-dependent interaction of ADP ribosylation factor 1 with coatomer subunit beta. Proc Natl Acad Sci 94: 4418–4423 [PMC free article] [PubMed] [Google Scholar]
  • Zhao L, Helms JB, Brunner J, Wieland FT 1999. GTP-dependent binding of ADP-ribosylation factor to coatomer in close proximity to the binding site for dilysine retrieval motifs and p23. J Biol Chem 274: 14198–14203 [PubMed] [Google Scholar]
  • Zhao X, Lasell TK, Melancon P 2002. Localization of large ADP-ribosylation factor-guanine nucleotide exchange factors to different Golgi compartments: Evidence for distinct functions in protein traffic. Mol Biol Cell 13: 119–133 [PMC free article] [PubMed] [Google Scholar]
  • Zhao X, Claude A, Chun J, Shields DJ, Presley JF, Melancon P 2006. GBF1, a cis-Golgi and VTCs-localized ARF-GEF, is implicated in ER-to-Golgi protein traffic. J Cell Sci 119: 3743–3753 [PubMed] [Google Scholar]
  • Zink S, Wenzel D, Wurm CA, Schmitt HD 2009. A link between ER tethering and COP-I vesicle uncoating. Dev Cell 17: 403–416 [PubMed] [Google Scholar]

Articles from Cold Spring Harbor Perspectives in Biology are provided here courtesy of Cold Spring Harbor Laboratory Press

-