CHAPTER SIX

Formation and Maintenance of the Golgi Apparatus in Plant Cells Yoko Ito*, Tomohiro Uemura*, Akihiko Nakano*,†,1

*Department of Biological Sciences, Graduate School of Science, University of Tokyo, Bunkyo-ku, Tokyo, Japan † Live Cell Molecular Imaging Research Team, RIKEN Center for Advanced Photonics, Wako, Saitama, Japan 1 Corresponding author: e-mail address: [email protected]

Contents 1. Introduction 2. Features of Plant Golgi as Compared to Animal and Yeast Cells 2.1 Morphology 2.2 Dynamics 2.3 Functions 3. ER–Golgi Transport 3.1 COPI vesicle 3.2 COPII vesicle and ERES 3.3 ERGIC 4. Intra-Golgi Trafficking 4.1 Models of intra-Golgi trafficking 4.2 Tethering and fusion of intra-Golgi COPI vesicles 5. Formation and Maintenance of Stacked Structure 5.1 Golgi apparatus as stacked cisternae 5.2 Search for the “Golgi stacking factors” 5.3 Golgi biogenesis upon mitosis 5.4 Golgi regeneration after BFA treatment 6. Relationship Between the Golgi Apparatus and the TGN 7. Concluding Remarks Acknowledgments References

222 222 222 227 231 235 235 242 252 253 253 256 258 258 259 261 263 265 267 267 268

Abstract The Golgi apparatus plays essential roles in intracellular trafficking, protein and lipid modification, and polysaccharide synthesis in eukaryotic cells. It is well known for its unique stacked structure, which is conserved among most eukaryotes. However, the mechanisms of biogenesis and maintenance of the structure, which are deeply related to ER–Golgi and intra-Golgi transport systems, have long been mysterious. Now having

International Review of Cell and Molecular Biology, Volume 310 ISSN 1937-6448 http://dx.doi.org/10.1016/B978-0-12-800180-6.00006-2

#

2014 Elsevier Inc. All rights reserved.

221

222

Yoko Ito et al.

extremely powerful microscopic technologies developed for live-cell imaging, the plant Golgi apparatus provides an ideal system to resolve the question. The plant Golgi apparatus has unique features that are not conserved in other kingdoms, which will also give new insights into the Golgi functions in plant life. In this review, we will summarize the features of the plant Golgi apparatus and transport mechanisms around it, with a focus on recent advances in Golgi biogenesis by live imaging of plants cells.

1. INTRODUCTION The Golgi apparatus is a single-membrane-bounded organelle in eukaryotic cells, and its pivotal roles in intracellular membrane trafficking are well established since its first description. The Golgi apparatus consists of several flattened membrane sacs called cisternae, which are stacked in layers like pancakes. In each Golgi stack, cisternae are polarized between the cis side, receiving cargo from the endoplasmic reticulum (ER), and the trans side, sending cargo forward to post-Golgi organelles (Mellman and Warren, 2000). Along this axis, many glycosylation enzymes are arranged as a gradient, so that cargo proteins can be modified sequentially by passing across the Golgi from cis to trans (Dunphy and Rothman, 1985; Moore et al., 1991). This stacked structure, which makes the Golgi apparatus “one of the most easily identified components of the cytoplasm” (Porter, 1961), is the biggest and unique morphological feature of this organelle. Although it has been attracting many biologists for a long time, little is known about how this structure is formed and maintained. In this review, we summarize the knowledge about the plant Golgi morphology, dynamics, functions, and underlying molecular mechanisms in comparison with yeast and animals. In addition, we highlight the recent data by live imaging approaches to reveal the biogenesis process of the stacked structure and its relationship with the ER–Golgi trafficking.

2. FEATURES OF PLANT GOLGI AS COMPARED TO ANIMAL AND YEAST CELLS 2.1. Morphology When Camillo Golgi first reported about the peculiar structure he had observed in the Purkinje cells of the cerebellum, he designated the structure as “apparato reticolare interno (internal reticular apparatus)” with fine and

Golgi Apparatus of Plant Cells

223

elegant network (Golgi, 1898). He also mentioned in his later writing that the reticular apparatus occupied the perinuclear zones and the nucleus was somehow embraced by it (Bentivoglio, 1998). However, these morphological features of the “classical Golgi” were soon questioned by the researchers who tried to identify the Golgi apparatus in invertebrate cells. When they applied Golgi’s staining method to invertebrate cells like insects or molluscs, discrete structures (granules, curved rods, and filaments) were observed (Beams and King, 1932). At that time, because of the absence of electron microscopic techniques, the existence of the Golgi apparatus itself was still under intense debate even in mammalian cells (Beams and Kessel, 1968; Bentivoglio, 1998). Moreover, since the appearance of the invertebrate Golgi apparatus was quite different from that of mammalian nerve cells, their homology drew much controversy (Beams and King, 1932). In the midst of the debates, it was suggested to call the discrete Golgi materials in insect cells as “dictyosomes,” the term that had been originally used to describe the pieces into which the Golgi apparatus fragmented during mitosis in vertebrate cells (Bowen, 1920). As regards the Golgi apparatus in plants, among suggestions that put the “vacuolar system” or the plastids as the homologs of the Golgi (Beams and Kessel, 1968), Bowen found that there were intracellular structures resembling the dictyosomes in many kinds of plant cells. They were independent from all the cytological structures previously described and he called them “osmiophilic platelets” (Bowen, 1928). In addition to the similarity in their appearance, it was demonstrated that the “osmiophilic platelets” occupied the same relative position as did the Golgi apparatus in animal cells upon ultracentrifugation (Beams and King, 1935a,b). In 1951, the first application of electron microscopy to the Golgi apparatus finally confirmed its existence (Dalton, 1951). Soon after, with the development of techniques for electron microscopic observation, the fine structure of the Golgi apparatus in vertebrate cells was revealed: concentrically arranged numerous flattened sacs of smooth-surfaced membranes with small vesicles (Dalton and Felix, 1954; Sjo¨strand and Hanzon, 1954). This characteristic structure of the Golgi apparatus was also found in insect cells (Beams et al., 1956), and of course, in various kinds of plant cells (Beams and Kessel, 1968). After all the half-century of discussion, the Golgi apparatus became to be recognized as an intracellular structure, which is conserved among most eukaryotic cells. Today, the term “osmiophilic platelets” is no longer in use, however, “dictyosome” still remains as a word to represent the Golgi stacks especially in plant sciences.

224

Yoko Ito et al.

As we can see from the history of the Golgi discovery, in spite of the conservation of their fine stacked structure at the electron microscopic level, the morphology of the whole Golgi apparatus is quite different among species. The plant Golgi apparatus consists of numerous, sometimes many hundreds of individual stacks of disk-like cisternae, and they are usually dispersed throughout the cytoplasm (Fig. 6.1A; Dupree and Sherrier, 1998; Hawes et al., 2008b). Each unit has a clearly stacked structure and cis–trans polarity, which can be confirmed either by the fine architecture under the electron microscope or by the localization of the marker proteins such as glycosylation enzymes and proteins involved in the trafficking machinery (Dupree and Sherrier, 1998; Ito et al., 2012; Schoberer and Strasser, 2011; Staehelin and Kang, 2008). This plant-style Golgi organization is also observed in protozoa, fungi (such as Schizosaccharomyces pombe and Pichia pastoris), and invertebrates (Wei and Seemann, 2010). On the other hand in vertebrate cells, as Camillo Golgi observed, the Golgi apparatus typically exhibits a twisted network, which is called the “Golgi ribbon” after its appearance under the light microscope

Figure 6.1 Comparison of the Golgi organization among eukaryotes. (A) In plant cells, individual Golgi stacks with clearly stacked structure are spread in the cytoplasm. Actin filaments are often arranged along the ER, and the Golgi stacks move on them by myosin motors. (B) In mammalian cells, microtubules are organized in a radial pattern around the centrosome. The Golgi stacks are centralized by dynein motors on the microtubules and interconnected to form the Golgi ribbon. ERGIC is formed near the ER, and ERGIC itself or transport vesicles from it travel to the Golgi ribbon along the microtubules. (C) In S. cerevisiae, the Golgi apparatus rarely forms stacks, and individual cis, medial, and trans cisternae scatter in the cytoplasm.

Golgi Apparatus of Plant Cells

225

(Fig. 6.1B; Lowe, 2011; Wei and Seemann, 2010). The Golgi ribbon is closely associated with the centrosome (or the microtubule-organizing center; MTOC), which is usually located near the nucleus in nonpolarized mammalian cells (Rios and Bornens, 2003). Because the Golgi ribbon is fragmented into many pieces with typical stacked organization (the ministacks) by depolymerizing microtubules (Polishchuk et al., 1999), it can be thought that the Golgi ribbon is an interconnected assemblage of the ministacks, which are very similar to the “dictyosome”-style Golgi in nonvertebrate organisms. Moreover, even in mammals, the Golgi apparatus naturally exhibits isolated stacks in certain types of cells (such as oocytes and skeletal muscle cells; Matteis et al., 2008). Hence, each discrete stacked structure as a minimum unit of the Golgi is thought to be fundamentally conserved among eukaryotes. However, there are exceptions. In the budding yeast Saccharomyces cerevisiae, the Golgi membranes rarely form stacks and scatter in the cytoplasm as individual cisternae, which exhibit tubular network-like structure. Each cisterna can be defined as cis, medial, and trans by the localization of the marker proteins (Fig. 6.1C; Matsuura-Tokita et al., 2006; Preuss et al., 1992; Rambourg et al., 1993). Moreover, in certain parasitic protists, the stacked Golgi structure was never observed and thus these species were considered as “Golgi-lacking” organisms ( Je´kely, 2008; Mowbrey and Dacks, 2009). It is now thought that species without Golgi stacks such as S. cerevisiae have evolved from the ancestors with stacked Golgi according to phylogenetic analyses (see Section 5.1; Mowbrey and Dacks, 2009). Although major progress has been made in the Golgi research by using mammalian cells, the complexity of the Golgi ribbon makes it difficult to observe the stacked structure in living cells by light microscopy. S. cerevisiae greatly contributed to resolve the molecular mechanisms of intracellular trafficking as a model organism (Schekman, 1992), and moreover, Ben Glick’s group and ours took advantage of the unstacked feature of the Saccharomyces Golgi to demonstrate the cisternal maturation (see Section 4.1; Losev et al., 2006; Matsuura-Tokita et al., 2006). However, we cannot investigate cisternal stacking in S. cerevisiae. By contrast, thanks to the simple and elegant structure, the Golgi stacking can be easily studied by light microscopy in plant cells. By introducing multiple marker proteins, each of which has distinctive intra-Golgi localization and a different colored fluorescent tag, the plant Golgi stacks are visualized as many punctate structures with the markers localized in layers (Fig. 6.2). On top of this, even though the size of the plant Golgi stacks can vary between species, cell types,

226

Yoko Ito et al.

Figure 6.2 Visualization of the plant Golgi stacks in living cells. (A) Confocal image of a tobacco BY-2 cell expressing GFP-AtSYP31 (cis-Golgi, green) and ST-mRFP (trans-Golgi, magenta). Arrowheads indicate the Golgi stack used for the analysis in B. Scale bar, 10 mm. (B) The fluorescence profile along the arrow across a Golgi stack.

and cell conditions, they are generally larger than the counterparts in other multicellular model organisms with dispersed Golgi [in diameter under electron microscopes, 370 nm in tissue culture S2 cells of Drosophila melanogaster (Kondylis and Rabouille, 2009), 200 nm in body-wall muscles and 500–800 nm in macrophage-like coelomocytes in Caenorhabditis elegans (Luo et al., 2011), 800 nm in tobacco BY-2 cultured cells (Nebenfu¨hr et al., 1999)]. The morphological features of the plant Golgi provide a great advantage for detailed analysis of the stacked structure by live imaging.

Golgi Apparatus of Plant Cells

227

2.2. Dynamics In plant cells, the majority of organelle movement is actomyosin dependent (Sparkes, 2011). Owing to the development of live imaging techniques, it has become clear that the Golgi apparatus is not an exception; the Golgi movement is greatly inhibited by actin disrupting drugs such as cytochalasin D and Latrunculin A, and by the myosin ATPase inhibitor 2,3-butanedione monoxime (Boevink et al., 1998; Nebenfu¨hr et al., 1999). In addition, by confocal microscopy, movement of the Golgi stacks can be observed along the actin filaments in living cells (Fig. 6.3). From the observation in tobacco BY-2 cells whose Golgi stacks are visualized by cis- and trans-Golgi markers tagged by different colors, the stacks are suggested to have a slight preference to direct the cis side ahead during the movement along the actin filaments, although the trans side can also

Figure 6.3 Plant Golgi stacks move along actin filaments. Time-lapse observation of a tobacco BY-2 cell expressing GFP-AtSYP31 (cis-Golgi, green), ST-mRFP (trans-Golgi, red), and Lifeact-Venus (actin filaments, blue). The Golgi stack indicated by the arrowheads moves toward the right. The indicated times mean the elapsed time. Scale bar, 5 mm.

228

Yoko Ito et al.

go forward and the stacks sometimes roll to change the leading side (Madison and Nebenfu¨hr, 2011). Microtubules, the other major cytoskeletal system, is not required for long-range movement of the plant Golgi stacks, because the treatment by microtubule disrupting drugs such as nocodazole, colchicine, and propyzamide does not have significant effects (Nebenfu¨hr et al., 1999). Plant myosins are classified into classes VIII and XI; the latter is structurally similar to class V myosins of metazoa and fungi, and is thought to share the common function in organelle movement (Avisar et al., 2008; Sparkes, 2011; Tominaga and Nakano, 2012). Several class XI myosins are shown to have overlapping functions and contribute to the Golgi movement to various degrees from knockout analyses, RNAi, and dominant negative inhibition in Arabidopsis and tobacco cells (Tominaga and Nakano, 2012). Observation of the Golgi movement in Arabidopsis knockout lines has recently suggested that myosin XI-K is the main contributor among them (Avisar et al., 2012). The dynamic flow of the cytoplasm including many organelles in plant cells has been long known as the cytoplasmic streaming (Shimmen and Yokota, 2004). Recent work has suggested that the unidirectional stream of the ER meshwork, which is driven primarily by myosin XI-K, hauls the cytosol and thus causes the cytoplasmic streaming (Ueda et al., 2010). The plant Golgi stacks may participate in this passive cytoplasmic streaming as well; however, their movement is not simple. The Golgi stacks exhibit stop-and-go motions; they alternate between pause for seconds to minutes and rapid directed movement, for example, with the velocity up to 7 mm/s in highly vacuolated Arabidopsis root cells (Akkerman et al., 2011; Boevink et al., 1998; Nebenfu¨hr et al., 1999). It is suggested that this Golgi motility pattern depends on the actin configuration (Akkerman et al., 2011), and that cortical microtubules establish “landmark sites,” which cause frequent pause of organelles not only the Golgi apparatus but also RNA processing bodies, peroxisomes, mitochondria, and the tubular ER junctions (Hamada et al., 2012), but the functional significance of such features of cytoskeletons is not clear. The Golgi motions were initially related to ER–Golgi transport; the Golgi stacks were proposed to receive cargo from the ERES (ER exit/export site) during stationary phases, because the Golgi stacks traveling in the same track appeared to stop occasionally at the same positions (Nebenfu¨hr and Staehelin, 2001; Nebenfu¨hr et al., 1999). This model was supported by the result that the inhibition of ER–Golgi transport

Golgi Apparatus of Plant Cells

229

by overexpression of a dominant negative form of RABD2A reduced the Golgi movement (Batoko et al., 2000). However, evidence to argue against this model was reported by fluorescence recovery after photobleaching (FRAP) analysis, which showed that ER-to-Golgi protein transport occurred while the Golgi stacks were moving (daSilva et al., 2004). This question goes together with the debate about the relationship between the Golgi stacks and ERES in plant cells (see Section 3.2.4), and is still under controversy. In addition, cellulose synthase complexes are suggested to be inserted to the plasma membrane from the Golgi, when the Golgi stacks pause on cortical microtubules (Crowell et al., 2009; Gutierrez et al., 2009). More experimental data are awaited, which visualize the Golgi stacks and their cargo transport. In contrast to plant Golgi, which uses actin filaments as their highways (Figs. 6.1A and 6.3), mammalian Golgi uses microtubules for long-range transport, although actin filaments also take parts in the Golgi structural organization and carrier vesicle formation (Egea et al., 2013). Mammalian cells have radial organization of microtubules around the centrosome with minus-ends anchored at the centrosome and plus-ends extending to the cell periphery (Fig. 6.1B; Egea and Rı´os, 2008). The minus-end-directed motor proteins, mainly dynein-1, collect the Golgi stacks along the microtubules and form the Golgi ribbon (Brownhill et al., 2009). A common feature of the cell types that do not have a continuous Golgi ribbon (from plant cells to mammalian oocytes and myotubes) is the absence of the radial patterned microtubule around the MTOC, indicating that the microtubule organization plays important roles in Golgi ribbon formation (Matteis et al., 2008). In addition, mammalian Golgi itself has been reported to act as an MTOC to nucleate microtubules both in vitro and in vivo, and the Golgi stacks can selforganize into a single continuous unit using self-derived microtubules in cytoplasts without centrosomes (Su¨tterlin and Colanzi, 2010; Wei and Seemann, 2009). Mammalian Golgi and the centrosome–microtubule system functionally interact with each other, not only in interphase but also during mitosis, and such relationship is not found in organisms with dispersed Golgi stacks (Su¨tterlin and Colanzi, 2010). Mammalian Golgi ribbon is relatively stable, but both the Golgi and the centrosome reorient to the traveling direction in many cell types upon migration (Brownhill et al., 2009). Moreover, dispersion or mislocalization of the Golgi ribbon without affecting centrosome-derived microtubule organization perturbs directional cell migration, while normal

230

Yoko Ito et al.

secretion ability is not affected (Hurtado et al., 2011; Yadav et al., 2009). The reason why mammalian Golgi apparatus forms a ribbon and behaves as a single organelle is still unclear; however, from the facts above, it is widely thought that Golgi ribbon formation maximizes polarized secretion efficiency. Then, are the cells without Golgi ribbon unable to carry out polarized secretion? Of course they can, in alternative ways. In S. cerevisiae, the Golgi cisternae cluster just below the growing bud during the initiation of new daughter cells, and they are also concentrated near the site of cell wall synthesis during cytokinesis (Preuss et al., 1992). This Golgi clustering can be interpreted as transient Golgi ribbon formation near the secretion target, while the cisternae are not interconnected. In D. melanogaster, the dispersed Golgi stacks are almost always found in close association with ERES (Kondylis and Rabouille, 2009), and a thousand of the ERES–Golgi units are evenly distributed in oocytes (Herpers and Rabouille, 2004). Among them, only the ERES–Golgi units in the dorsal/anterior corner transport Gurken protein, whose luminal fragment is secreted locally at the dorsal/anterior corner to specify the anteroposterior and dorsoventral axes, whereas all the ERES–Golgi units in the cell are equal in transport ability. This polarized transport is suggested to depend on polarized localization of gurken mRNA combined with its local translation (Herpers and Rabouille, 2004). This indicates that the dispersed Golgi stacks could functionally differentiate to transport different proteins. In support of this, the subset of Golgi stacks are proposed to possess different glycosylation enzymes in a single Drosophila imaginal disk cell (Yano et al., 2005). Also in plant cells, the Golgi stacks are shown to localize in the region several microns behind the growing tips of root hairs and pollen tubes, while the extreme tips are occupied by vesicles including Golgi-derived secretory vesicles (Campanoni and Blatt, 2007; Dupree and Sherrier, 1998). This Golgi behavior might be similar to the Golgi clustering in S. cerevisiae for polarized secretion, but the association of the Golgi stacks with growing regions is not absolute in plant cells (Dupree and Sherrier, 1998). It is unclear whether the Golgi stacks are actively gathered, or their localization is just the consequence of the concentration of the cytoplasm and only the vesicles are targeted to the growing tips. Another major event that requires polarized secretion in plant cells is cell plate formation upon cytokinesis. It is well known that the Golgi–TGNderived vesicles carrying cell wall components are targeted to the division

Golgi Apparatus of Plant Cells

231

plane, and the cell plate is formed by fusion of these vesicles ( Ju¨rgens, 2004; Staehelin and Hepler, 1996). In the past without live imaging techniques, the plant Golgi stacks were thought to have no particular organization during mitosis (Hepler and Wolniak, 1984). However, live imaging observation in tobacco BY-2 cells revealed that the general Golgi streaming stops when the cells enter mitosis, and a subset of stacks begin to redistribute around the perinuclear region. Later in metaphase, they aggregate around the mitotic spindle and in the equatorial region under the plasma membrane. This latter region accurately predicts the future site of cell division, and has been named the “Golgi belt.” During cytokinesis, many Golgi stacks localize around the phragmoplast, an array of microtubules required for proper targeting of the vesicles for the cell plate (Nebenfu¨hr et al., 2000). Such accumulation of the Golgi stacks has been also observed in Arabidopsis shoot apical meristem cells, in which up to 80% of the stacks localize in close proximity of the phragmoplast in early telophase (Seguı´-Simarro and Staehelin, 2006). This transition of the Golgi localization is thought to reflect the centralization of the secretory pathway for rapid cell plate synthesis. Although the orientation of the Golgi stacks and their associated TGN toward the target presumably affects polarized secretion efficiency, any biased orientation of the Golgi– TGN units has not been observed (Seguı´-Simarro and Staehelin, 2006). It is still unknown whether the plant Golgi stacks functionally differentiate in a single cell, like the case of ERES–Golgi units in Drosophila cells.

2.3. Functions 2.3.1 Subcompartmentalization of glycosylation enzymes The Golgi apparatus plays central roles not only in cargo transport and sorting but also acts as an important biosynthetic compartment where the complex oligosaccharides moieties of glycoproteins and glycolipids are assembled. This Golgi function as the biochemical “factory” is fundamentally conserved among eukaryotes, albeit the actual modification steps may differ between species or cell types (Gomord et al., 2010; Song et al., 2011). A widely accepted view on the sugar chain modification process is the assembly-line model, in which sequentially ordered glycosylation enzymes elongate glycan chains of substrates along their passage through the Golgi apparatus (Berger and Rohrer, 2008). Mammalian glycosidases and glycosyltransferases have been demonstrated to localize in the Golgi apparatus in specific order along the cis–trans axis with some overlap, and this distribution is thought to mirror the order of glycosylation steps (Igdoura et al.,

232

Yoko Ito et al.

1999; Nilsson et al., 1993; Rabouille et al., 1995). Altering the intra-Golgi localization of glycosyltransferases has been reported to affect glycosylation pattern of the secreted reporter glycoprotein (Grabenhorst and Conradt, 1999), which indicates that the ordered distribution of the enzymes contributes to correct glycosylation. The mechanisms to achieve such subcompartmentalization of the Golgi resident glycosylation enzymes have been intensively studied in mammalian and yeast cells. Most of the glycosylation enzymes are type II integral membrane proteins comprised of a short N-terminal cytoplasmic tail, a single transmembrane domain, a stem region, and a catalytic domain oriented into the Golgi lumen (Berger and Rohrer, 2008; Schoberer and Strasser, 2011; Tu and Banfield, 2010). The information of Golgi localization is in general located within the cytoplasmic tail, the transmembrane domain, and the stem region (CTS), while the localization mechanisms can vary from enzyme to enzyme (Tu and Banfield, 2010). For some of them, the length of the transmembrane domain seems to be important (Munro, 1995). For others, not the length but the composition of amino acids of the transmembrane domain, which promotes homo- or heterooligomerization, are significant for enzyme localization (Tu and Banfield, 2010). Such heterooligomerization of the glycosylation enzymes is also suggested to ensure sequential modification of the substrates (de Graffenried and Bertozzi, 2004). Still others, especially specific yeast glycosyltransferases, are suggested to have consensus sequences in the cytoplasmic tails to associate with other proteins that also bind to COPI coats, which promotes segregation of the enzymes into transport vesicles and contributes to their localization (Moremen et al., 2012; Schmitz et al., 2008; Tu and Banfield, 2010; Tu et al., 2008). Clearly, there is no single conserved way, and moreover, these hypothetical mechanisms can act in combination. This question about the Golgi enzyme subcompartmentalization mechanism cannot be independent from the big debate about intra-Golgi trafficking (see Section 4.1), and thus is still under controversy. Rat a-2,6-sialyltransferase (ST) expressed in plant cells exhibits the localization in the trans-Golgi cisternae, very similarly to its original host organism (Wee et al., 1998), and this localization is sufficiently accomplished only by its CTS region (Boevink et al., 1998). Thus, the mechanisms of localization of the glycosylation enzymes appear to be (at least partially) conserved between animals and plants. Indeed, the fluorescence-tagged version of the CTS region of rat ST is one of the most commonly used trans-Golgi membrane marker in plant cells now (Fig. 6.2). Furthermore, the sugar chain

Golgi Apparatus of Plant Cells

233

structure of glycoproteins can be manipulated by swapping the CTS regions of introduced glycosyltransferases with different enzymes, which presumably alter the localization of the chimera enzymes to that of the donor of the CTS region (Bakker et al., 2006; Strasser et al., 2009). Cloning of plant glycosylation enzymes began to catch up mammals and yeast (Boisson et al., 2001; Liebminger et al., 2009; Strasser et al., 1999, 2000, 2006, 2007), which has greatly facilitated studies on their intra-Golgi localization and trafficking by live imaging. Mechanisms to locate different enzymes are proposed similarly to mammals: depending on the length of transmembrane domain (Saint-Jore-Dupas et al., 2006) or homo- or heterodimerization through CTS regions (Schoberer et al., 2013). Although ER-to-Golgi export signals are found in the cytoplasmic tails of plant Golgi membrane proteins (Hanton et al., 2005; Schoberer et al., 2009; Yuasa et al., 2005), it is not clear whether they contribute to intra-Golgi localization.

2.3.2 N-linked glycosylation Among the posttranslational modifications on proteins, N-linked glycosylation, synthesis and processing of oligosaccharides on asparagine residues in Asn-X-Ser/Thr (X represents any amino acid residues) motifs, has been most extensively studied. In all eukaryotes, N-glycosylation is initiated in the ER by en bloc transfer of the preformed oligosaccharide (Glc3Man9GlcNAc2) to the nascent polypeptide chain during translation. Next, two glucose residues are cleaved off from the N-glycan by ER a-glucosidase I and II (GCSI and GCSII) to make Glc1Man9GlcNac2, which is recognized by the ER-resident lectin-like chaperones, calreticulin (CRT), and calnexin (CNX). Subsequently, the terminal glucose residue is trimmed and the glycoprotein is freed from CRT/CNX. However, if the protein is not properly folded, its N-glycan is reglucosylated and caught by CRT/CNX, resulting in retention in the ER or retrieval back from the Golgi. Such CRT/CNX-glycoprotein binding and liberation cycle continues until the protein attains its native structure (CRT/CNX cycle) (Caramelo and Parodi, 2008). Such function of N-glycosylation in the ER quality control (ERQC) is conserved among most eukaryotes. In Arabidopsis, null mutants of glycosylation enzymes in the early steps in N-glycosylation such as GCSI and GCSII are known to be lethal (Boisson et al., 2001; Burn et al., 2002). Recently, the N-glycan dependent ERQC is suggested to be required for proper function of EFR, a pattern recognition receptor in the plant immune system for the bacterial elongation

234

Yoko Ito et al.

factor-Tu (Farid et al., 2013; Liu and Howell, 2010; Saijo, 2010). EFR is thought to be affected by N-glycosylation in protein stability, accumulation, and ligand binding (Farid et al., 2013; Ha¨weker et al., 2010). Some other plant proteins have also been reported to need N-glycans for their catalytic activity, thermostability, subcellular localization, and secretion, implying contribution of N-glycan dependent ERQC to their proper structures (Ceriotti et al., 1998; Lige et al., 2001). After they clear the ERQC system, glycoproteins go forward to the Golgi apparatus, where their high-mannose-type N-glycans acquire further modifications to become complex-type N-glycans. In mammalian cells, complex N-glycans are thought to play many important roles. For example, deletion of N-acetylglucosaminyltransferase (GnTI), the glycosyltransferase thought to initiate N-glycan processing in the Golgi, causes deficiency in complex N-glycan formation and leads to embryonic lethality in mice (Ioffe and Stanley, 1994; Metzler et al., 1994). On the other hand, mice lacking a-mannosidase II (aM-II), an enzyme acts in the early stage of Golgi N-glycan processing just after GnTI, develop an autoimmune disease (Paulson, 2007). However, in spite of the conservation of N-glycan functions in the ER between animals and plants, very little is known about the role of N-glycan processing in the plant Golgi apparatus. It is because no physiological phenotype was reported for the mutants of Golgi N-glycan modification until recently. Arabidopsis mutants defective in the Golgi glycosylation enzymes including GnTI and aM-II cannot produce normal complex N-glycans, but exhibit no growth or morphological defect under normal environment (Strasser et al., 2004, 2006; von Schaewen et al., 1993). For EFR, which requires N-glycosylation in the ER, Golgi-type modifications are dispensable (Ha¨weker et al., 2010). Artificial suppression of complex N-glycan processing by RNAi of a1,3-fucosyltransferase (FucT) and b1,2-xylosyltransferase (XylT) in tobacco and duckweed does not show obvious phenotype (Cox et al., 2006; Strasser et al., 2008). In 2008, it was reported that A. thaliana mutants defective in Golgi N-glycosylation are highly sensitive to salt stress (Kang et al., 2008). Genetic analysis indicates that the function of plasma membrane-bound endo-b1,4-glucanse KORRIGAN 1 (KOR1), which is essential for cellulose biosynthesis of cell wall, depends on complex N-glycans. Moreover, the weak mutant of KOR1 is also salt sensitive. Therefore, complex N-glycans are suggested to be necessary for sufficient cell wall formation under salt stress (Kang et al., 2008). Further investigations especially about biotic and abiotic stress responses will provide more understanding of the functions of N-glycan processing in the plant Golgi apparatus.

Golgi Apparatus of Plant Cells

235

2.3.3 Cell-wall polysaccharide synthesis In addition to processing of sugar chains of glycoproteins, the Golgi apparatus in plant cells plays a crucial role in synthesis of noncellulosic cell wall polysaccharides. The primary cell wall of eudicots consists of cellulose microfibrils cross-linked by hemicellulose xyloglucan (XyG), and noncellulosic polysaccharides and proteins as a matrix filling around the XyG–cellulose network. Major noncellulosic polysaccharides in the primary wall except XyG are pectic polysaccharides; homogalacturonan (HG), rhamnogalacturonan I and II (RG-I and RG-II). Unlike the polysaccharide chains on glycoproteins initially synthesized in the ER, the noncellulosic cell wall polysaccharides are assembled exclusively in the Golgi apparatus (Driouich et al., 2012). Using antibodies recognizing either the XyG backbone or the side chains, XyG has been shown to appear in trans-Golgi and the TGN (Moore et al., 1991; Zhang and Staehelin, 1992). By a similar immunocytochemical approach, HG is suggested to be synthesized in unesterified form in cis and medial cisternae followed by methylesterification in medial and trans cisternae (Moore et al., 1991; Zhang and Staehelin, 1992). For RG-I, backbone structure is synthesized similarly to HG, and arabinose-containing side chains are supposed to be added in trans cisternae (Zhang and Staehelin, 1992). Information about localization and assembly of RG-II is still missing. Most of the enzymes involved in plant cell wall polysaccharide synthesis are type II membrane proteins, albeit some are multipass transmembrane proteins (Oikawa et al., 2013). Similarly to the N-glycan glycosyltransferases, the enzymes that add side chains to XyG have been shown to localize in different Golgi cisternae in the order they suggested to work in tobacco BY-2 clls (Chevalier et al., 2010). Furthermore, the understanding about protein complex formation has recently made significant advance by application of fluorescent imaging techniques such as bimolecular fluorescence complementation (BiFC), split luciferase complementation assay (SLCA), and Fo¨rster resonance energy transfer (FRET). Using these techniques, it is becoming clear that the protein–protein interaction of glycosylation enzymes to make complexes is crucial for cell wall polysaccharide synthesis as N-glycosylation (Oikawa et al., 2013).

3. ER–GOLGI TRANSPORT 3.1. COPI vesicle 3.1.1 COPI coat (coatomer) From cell-free reconstitution assays and cell fusion experiments to monitor the cargo transport between different mammalian Golgi complexes, original

236

Yoko Ito et al.

prediction was that intra-Golgi (especially anterograde) transport was achieved by vesicle mediated trafficking (Rothman, 1994). In 1986, de novo formation of nonclathrin-coated vesicles from isolated mammalian Golgi was shown. A subset of their coat proteins were identified and shown to make a complex (coatomer), and given the name coat proteins (COPs; Serafini et al., 1991; Waters et al., 1991). This “COP-coated vesicle” was subsequently renamed COPI vesicle when the existence of COPII vesicle was established (Barlowe et al., 1994). Now, it is known that coatomer is a heptameric complex consisting of a, b, b0 , g, d, e, and z subunits (Hsu et al., 2008; Popoff et al., 2011). They are classified into two subcomplexes; one complex composed of b-, g-, d-, and z-COP forms the inner core that binds cargo, and the other complex composed of a-, b0 -, and e-COP forms the outer lattice (Popoff et al., 2011). COPI machinery is well conserved among eukaryotes including plants. All the COPI components have been found to be encoded in the Arabidopsis genome (Bassham et al., 2008). Plant COPI vesicles were first visualized by immunoelectron microscopy in Arabidopsis and maize cells by antibodies against Arabidopsis g-COP and maize d- and e-COP (Pimpl et al., 2000). Using these antisera, plant coatomers have been detected on the cisternal rims of the Golgi apparatus and the vesicles including buds formed from the Golgi (Donohoe et al., 2007; Pimpl et al., 2000; Ritzenthaler et al., 2002). In the human genome, two distinct isoforms of g-COP (g1 and g2) and z-COP (z1 and z2) have been found (Blagitko et al., 1999; Futatsumori et al., 2000). These coatomer isoforms have been demonstrated to show different intra-Golgi localization, which indicates different budding sites or functions for each of them (Moelleken et al., 2007). Arabidopsis also codes multiple isoforms of COPI proteins except for g- and s-COP; two for a-, b-, e-COP, and three for b0 - and z-COP so far (Bassham et al., 2008; Robinson et al., 2007). Whether they localize and act differently is yet to be elucidated, but it has been reported by electron tomography that there are two types of COPI vesicles (COPIa and COPIb) in Arabidopsis and scale alga Scherffelia dubia according to their coat architecture, coat thickness, and cargo staining. Moreover, COPIa vesicles bud exclusively from cis-cisternae rims and occupy the space between cis-Golgi and the ER, while COPIb vesicles bud from medial- and trans-Golgi and are distributed around these cisternae. This may indicate that COPIa vesicles are involved in the transport between the ER and the Golgi, whereas COPIb vesicles function in intraGolgi transport (Donohoe et al., 2007, 2013).

Golgi Apparatus of Plant Cells

237

3.1.2 ARF1 COPI vesicle formation is regulated by a GTPase ADP-ribosylation factor 1 (ARF1). Like other small GTPases, ARF1 cycles between GDP-bound inactive cytosolic state and GTP-bound active membrane-anchored state. ARF1 has hydrophobic and myristoylated N-terminus, which is exposed and inserted into the lipid bilayer by a conformational change coupled with activation by nucleotide exchange (Antonny et al., 1997; Goldberg, 1998). Activation of ARF1 by GTP binding is necessary for coatomer recruitment on the Golgi membranes (Donaldson et al., 1992a). In plant cells, ARF1 localization was observed on the Golgi and COPI vesicle membranes by immunoelectron microscopy (Pimpl et al., 2000), and later the localization to the Golgi stacks was confirmed by immunofluorescence (Ritzenthaler et al., 2002) and live imaging of fluorescence-tagged ARF1 (Stefano et al., 2006; Xu and Scheres, 2005). It is noteworthy that the immunofluorescence and ARF1-GFP signals are observed as doughnut-like rings, which indicates that ARF1 localizes especially to the Golgi cisternal rims where COPI vesicles are supposed to bud (Ritzenthaler et al., 2002; Xu and Scheres, 2005). In addition to the Golgi stacks, it has been demonstrated that ARF1 localizes to the TGN by immunofluorescence (Paciorek et al., 2005; Tanaka et al., 2009) and ARF1-GFP (Tanaka et al., 2009; Xu and Scheres, 2005), and recently by immunoelectron microscopy (Stierhof and El Kasmi, 2010). Such ARF1 population is thought to regulate the formation of clathrin-coated vesicles from the TGN (Memon, 2004; Robinson et al., 2011). Surprisingly, it has been reported that barley ARF1 expressed in onion cells colocalizes with AtARA7, the marker of late endosomes/multivesicular bodies (MVBs; Bo¨hlenius et al., 2010). Plant TGN has another face as the early endosome (see Section 6), and the TGN is recently suggested to mature into MVB along the endocytic pathway. In this context, because the compartments with an intermediate nature between the TGN and MVB are observed when endocytosis is highly activated (Choi et al., 2013), the colocalization of barley ARF1 and AtARA7 might reflect such intermediate compartments. 3.1.3 ARF GEF and GAP For activation of ARF1, guanine nucleotide exchange factors (GEFs) are needed, and they are characterized by a conserved Sec7 catalytic domain ( Jackson and Casanova, 2000). Among mammalian ARF GEFs, GBF1 is proposed to regulate COPI recruitment on the cis-Golgi and BIGs are

238

Yoko Ito et al.

specialized for clathrin-coated vesicle formation at the TGN, while the others act at post-Golgi compartments (Manolea et al., 2008; Popoff et al., 2011). ARF1-GDP meets GBF1 on the membrane to be ARF1GTP, and recruits multiple effectors including coatomers, cargo proteins, lipid-modifying enzymes, and ARF GTPase-activating proteins (GAPs; Hsu et al., 2008; Popoff et al., 2011). Brefeldin A (BFA), an antibiotic compound produced by fungi, was known to cause rapid distribution of Golgi proteins into the ER and disappearance of the Golgi apparatus in mammalian cells (Lippincott-Schwartz et al., 1989). Subsequently, BFA was found to inhibit activation step of ARF1 (Donaldson et al., 1992b; Helms and Rothman, 1992). BFA is hypothesized to bind between ARF1-GDP and the Sec7 domain of the GEF and stabilize abortive ARF1-GDP:BFA:Sec7 domain complex, which leads to the inhibition of nucleotide exchange of ARF1 and finally dissociates ARF1 and COPI coat from the Golgi membranes (Chardin and McCormick, 1999). Among ARF GEF subgroups, only GBF and BIG types are found in plants (Teh and Moore, 2007). The first ARF GEF identified in plants was the Arabidopsis GBF protein GNOM (GN). Although animal and fungal GBF proteins act at the Golgi, GNOM was revealed to function mainly in the recycling process of auxin efflux carrier PIN1 from endosomes to the plasma membrane (Geldner et al., 2003). Later, GNOM-LIKE1 (GNL1), the closest homolog of GN in Arabidopsis, was identified and shown to be responsible for ER–Golgi trafficking. GNL1 primarily localizes to the Golgi stacks and regulates coatomer recruitment there (Richter et al., 2007). Because GN can functionally substitute for GNL1 but not vice versa, GN is supposed to gain the plant-specific function in addition to the ancestral role of GBF proteins in ER–Golgi trafficking (Richter et al., 2007). Since GNL1 is BFA resistant, BFA treatment does not cause Golgi disruption in Arabidopsis root cells unlike in mammalian and tobacco cells (Teh and Moore, 2007). In order to fuse with target membranes, coated vesicles need to be uncoated. For COPI vesicle uncoating, inactivation of ARF1 is necessary, which requires GAPs (Bigay et al., 2003; Melanc¸on et al., 1987; Reinhard et al., 2003). However, the ARF GAP activity was revealed necessary not only for vesicle uncoating but also for COPI cargo sorting and vesicle formation (Lee et al., 2005; Nie and Randazzo, 2006). The GAP activity of mammalian ARF GAP1 is accelerated by binding with coatomer, but this effect is inhibited by a p24 protein, which is suggested to act as a cargo

Golgi Apparatus of Plant Cells

239

receptor (Goldberg, 2000). Therefore, ARF GAP1 is presumed to act in “proofreading” of COPI vesicle formation, because only the coatomers with cargo proeins are thought to stay on the membrane long enough to be polymerized. Moreover, it has been also reported that ARF GAP1 activity increases as the membrane curvature approaches that of a typical transport vesicle, which might ensure vesicle uncoating after completion of vesicle formation (Bigay et al., 2003). In plant cells, although the Arabidopsis genome encodes 15 ARF GAP domain (AGD) proteins, most of their functions are unknown. AGD7, the closest homolog of human ARF GAP1, has been revealed to localize to the Golgi apparatus and regulate ARF1 activity in ER–Golgi trafficking (Min et al., 2007). In addition, AGD8 and AGD9, homologs of Glo3p-type ARF GAPs functioning in COPI vesicle formation in animal and yeast cells, are reported to recruit ARF1-GDP from the cytosol to the Golgi membranes (Min et al., 2013). Such a role has not been reported for mammalian or yeast ARF GAPs. 3.1.4 Golgi-to-ER transport Although COPI vesicles were thought to act in the anterograde trafficking in the early studies, this initial view was changed by subsequent findings indicating their roles in Golgi-to-ER retrograde transport. Soluble ER proteins have a common KDEL or HDEL motif in their C-termini in animal and yeast cells, and this motif is recognized by the receptor coded by yeast ERD2 gene and its homologs in animals. Erd2 captures leaked ER luminal proteins mainly at the cis-Golgi or ER-Golgi intermediate compartment (ERGIC), and returns them to the ER (Lewis and Pelham, 1990; Lewis et al., 1990; Semenza et al., 1990). This Golgi-to-ER transport step has been shown to depend on COPI vesicles (Girod et al., 1999; Lewis and Pelham, 1996). For transmembrane proteins, ER retrieval signals in the cytosolic domain are known to bind directly to coatomer subunits (Letourneur et al., 1994; Michelsen et al., 2007). In yeast cells, a subset of ER membrane proteins are recognized through their transmembrane region by Rer1 at the Golgi and retrieved to the ER, and this process also depends on COPI vesicles (Sato et al., 1995, 2001, 2003). These recognition systems for Golgi-to-ER retrograde traffic by COPI vesicles are also conserved in plants. C-terminal KDEL and HDEL sequences are sufficient for ER localization of soluble proteins (Denecke et al., 1992). The Arabidopsis homolog of ERD2 can complement the yeast erd2 deletion mutant (Lee et al., 1993), and localizes to the Golgi stacks in

240

Yoko Ito et al.

Arabidopsis and tobacco cells (Boevink et al., 1998; Saint-Jore et al., 2002; Takeuchi et al., 2000). There are three RER1 homologs in Arabidopsis (RER1A, RER1B, and RER1C) and all can complement yeast rer1 deletion in different degrees with RER1B being the best (Sato et al., 1999). GFPRER1B has been demonstrated to localize to the cis-Golgi in Arabidopsis and tobacco cultured cells (Ito et al., 2012; Takeuchi et al., 2000). ER-localized proteins with the COPI-interacting ER retrieval signals have been also identified (Benghezal et al., 2000; Boulaflous et al., 2009; Contreras et al., 2004; McCartney et al., 2004; Montesinos et al., 2012). Because BFA inhibits the formation of COPI vesicles responsible for Golgi-to-ER transport, it seems puzzling that the secretion is disrupted and the Golgi enzymes relocalize to the ER by BFA treatment in many cell types. In yeast, mutation in UFE1, which encodes the SNARE protein necessary for COPI vesicle fusion with the ER, exhibits blockade of secretion and ER retention of a vacuolar protein (Lewis and Pelham, 1996). In plant cells, overexpression of GDP-locked dominant negative form of ARF1 (ARF1T31N) causes absorbance of ERD2 and CTS regions of ST and XylT into the ER, and inhibits secretion of a-amylase (Stefano et al., 2006; Takeuchi et al., 2002). Moreover, GTP-locked constitutive active ARF1 mutant protein (ARF1Q71L) causes accumulation of ERD2 and a vacuolar protein sporamin in the ER (Takeuchi et al., 2002). Because weak expression of this construct only inhibits Golgi-to-ER transport (Langhans et al., 2008), blockade of ER export is supposed to be a secondary effect. From these facts, the depression of ER-to-Golgi anterograde trafficking by BFA treatment is thought to be the indirect consequence of the inhibition of the retrograde trafficking, which is presumably necessary for salvaging components for COPII vesicle formation (such as SNAREs). However, the disappearance of the Golgi apparatus cannot be explained only by disruption of transport between the ER and the Golgi. Actually in mammalian cells, it has been reported that the Golgi apparatus does not disappear by combining a cationic ionophore monensin with BFA, in spite of the dissociation of ARF1 and coatomer from the Golgi membranes and the blockade of ER export (Barzilay et al., 2005). In tobacco BY-2 cells, hybrid structures of the ER and the Golgi stacks have been observed upon BFA treatment by electron microscopy, which suggested direct fusion of the two organelles (Ritzenthaler et al., 2002). The absorption of the Golgi proteins to the ER was proposed to result from this fusion. It was hypothesized that this phenomenon was due to Golgi accumulation of the SNARE proteins responsible for fusion of Golgi-derived COPI vesicles to the ER, although there is no experimental support. Anyway, these data indicate that the

Golgi Apparatus of Plant Cells

241

existence of the Golgi apparatus depends on the delicate balance between anterograde and retrograde trafficking between the ER and the Golgi. COPI vesicles are also believed to be responsible for intra-Golgi trafficking. About this transport pathway, we will describe in Section 4. 3.1.5 Tethering factors and SNAREs The fundamental process of membrane fusion is mediated by long-range tethering factors mostly regulated by RAB GTPases, and short-range SNARE proteins (Saito and Ueda, 2009; Szul and Sztul, 2011). For the Golgi-to-ER COPI vesicles, tethering to the ER membrane is mediated by Dsl1 trimeric complex in yeast cells (Ren et al., 2009). Dsl1 complex is composed of Dsl1, Tip20, and Dsl3/Sec39, and Dsl1 binds to a- and d-COP (Andag and Schmitt, 2003; Reilly et al., 2001). Dsl1 complex stably localizes to the ER membrane independently to any RAB GTPases (Sztul and Lupashin, 2009). Membrane fusion is mediated by ER-localized Q-SNAREs (Qa: Ufe1, Qb: Sec20, and Qc: Use1) and R-SNARE Sec22 (Ballensiefen et al., 1998; Burri et al., 2003; Dilcher et al., 2003; Lewis and Pelham, 1996). Dsl1 complex interacts with the SNARE complex and is suggested to promote SNARE complex assembly during membrane fusion (Kraynack et al., 2005; Ren et al., 2009). In mammalian cells, the homologous complex of Dsl1 composed of ZW10, RINT-1, and NAG has been identified and also interacts with ER Q-SNAREs (Schmitt, 2010; Sztul and Lupashin, 2009). Among three subunits of Dsl1 complex, a homolog of Tip20 was first identified in Arabidopsis. It was isolated as the gene responsible for maigo2 (mag2) mutant, which accumulates seed storage proteins in the ER. MAG2/ AtTIP20 protein is associated with the ER membrane, and interacts with SNAREs AtSEC20 and AtSYP81/AtUFE1 (Li et al., 2006). CFP-tagged MAG2/AtTIP20 expressed in tobacco leaf epidermal cells has been shown to colocalize with the Golgi stacks and move together (Lerich et al., 2012). A recent study identified three MAG2/AtTIP20 interators designated MAG2 INTERACTING PROTEINS (MIPs,) and the mutant plants lacking any one of them developed abnormal structures with storage proteins in the ER of seed cells. Because MIP1 and MIP2 were revealed to have a ZW10 domain and a Sec39 domain, respectively, the MAG2/AtTIP20– MIP1–MIP2 complex is suggested to be homologous to Dsl1 complex. However, this complex has the fourth subunit, MIP3, which contains a Sec1 domain. This is the first report of a Sec1-containing protein involved in ER–Golgi trafficking including animals and yeast (Li et al., 2013).

242

Yoko Ito et al.

By live-cell imaging using GFP-tagged proteins, six SNAREs are shown to localize to the ER in plant cells: Qa-SNARE AtSYP81/AtUFE1; Qc-SNARE AtSYP71, AtSYP72, and AtSYP73; and R-SNARE AtSEC22 and AtVAMP723 (Uemura et al., 2004). No Qb-SNARE was found on the ER membrane in this study, but AtSEC20 is suggested to localize to the ER. In addition, AtUSE1 was recently reported to localize to the ER (Lerich et al., 2012). Since overexpression of AtSYP81/AtUFE1 causes relocalization of Golgi enzymes, it is thought to function in the transport between the ER and the Golgi (Bubeck et al., 2008). SYP7 family is not conserved in yeast and animals, and its involvement in COPI vesicle transport is not clear (Sanderfoot et al., 2000). However, AtSYP72 tagged by fluorescent protein marks punctate signals along the ER tubules, and AtSYP81/AtUFE1 and COPI dependent Golgi-to-ER cargo proteins colocalize at the puncta only when coexpressed with AtSYP72. Such punctate signals do not move with the Golgi stacks, but when the Golgi stacks are immobilized by actin depolymerization, they colocalize with the Golgi signals. Therefore, it is proposed that Golgi-to-ER retrograde vesicles are tethered by Dsl1 complex while the Golgi stack is moving, and fuse to the ER during temporary pauses of the Golgi on the AtSYP72 puncta (Lerich et al., 2012).

3.2. COPII vesicle and ERES 3.2.1 COPII vesicle formation COPII coat assembly in S. cerevisiae is initiated by activating small GTPase Sar1 by its GEF, Sec12 (Barlowe and Schekman, 1993; Nakano and Muramatsu, 1989). Sec12 has been revealed to interact with Rer1 through its transmembrane domain, therefore, Sec12 is regulated to localize to the ER membrane by Golgi-to-ER COPI vesicles, which results in restriction of Sar1 activation on the ER (Sato et al., 1996). Sar1 has a hydrophobic a-helix at the N-terminus similarly to one of its closest evolutionary relatives Arf1, and is believed to associate with the membrane by embedding the N-terminal domain into the bilayer upon activation (Bi et al., 2002; Huang et al., 2001). In contrast to heptameric coatomer, which is recruited to the membrane en bloc, COPII coat complexes (Sec23/24 and Sec13/31) are recruited sequentially to form layers. Activated Sar1 recruits Sec23/24 complex first by interacting with Sec23 subunit to form Sec23/24-Sar1 “prebidding complex” (Bi et al., 2002; Yoshihisa et al., 1993). The prebudding complexes recognize cargo proteins and compose the inner layer of COPII coat (Kuehn et al., 1998). The assembly of the prebudding

Golgi Apparatus of Plant Cells

243

complex is stabilized through interactions with transmembrane cargo proteins, and different kinetics between the disassembly of the prebudding complex and the cycle of Sar1-GTP hydrolysis facilitates proper and efficient cargo sorting and vesicle formation (Sato and Nakano, 2005). The outer layer is composed of Sec13/31 complex organized into a cage (Matsuoka et al., 2001). Sec13/31 complex interacts with the prebudding complex through binding of Sec31 to Sec23 and Sar1 (Bi et al., 2007). The outer cage assembly is likely to collect the prebudding complexes and drive membrane deformation. COPII components are highly conserved in eukaryotes. Notably, the Arabidopsis genome encodes multiple isoforms for each COPII component (two Sec12, five Sar1, seven Sec23, three Sec24, two Sec31, and two Sec31; Robinson et al., 2007). Among them, mutant plants have been reported only for Sec24 isoforms. A recessive missense point mutation in AtSEC24A, which causes partial retention of Golgi and secretory proteins in the abnormally shaped ER, has been independently identified by two groups by forward genetics combined with visual screening using Golgi or ER markers. Moreover, Atsec24A null mutant was revealed to be lethal, indicating that AtSEC24A is essential (Faso et al., 2009; Nakano et al., 2009). Recently, mutants of AtSEC24B and AtSec24C were reported, although the organelles and intracellular transport is yet to be investigated in detail (see Section 3.2.2; Tanaka et al., 2013). The functions of the plant homologs of Sec12 and Sar1 have been confirmed in alternative ways. AtSEC12 suppresses the yeast sec12 null mutant (d’Enfert et al., 1992). In addition, overexpression of AtSEC12 in tobacco protoplasts inhibits secretion, and this effect is rescued by additional overexpression of AtSAR1B (Phillipson et al., 2001). AtSAR1B and tobacco (Nicotiana tabacum) SAR1 also have been revealed to suppress yeast sar1 null mutant and temperature-sensitive sec12 or sec16 mutants, which indicates the functional conservation (d’Enfert et al., 1992; Takeuchi et al., 1998). The GTP-locked dominant mutant protein of AtSAR1B causes retention of Golgi and vacuolar proteins in the ER and inhibits secretion (Phillipson et al., 2001; Takeuchi et al., 2000). GTP-locked AtSAR1A is also reported to inhibit secretion but to a less extent than AtSAR1B, and the membrane binding affinity of AtSAR1B is stronger than AtSAR1A (Hanton et al., 2008). Whether such differences reflect functional difference is not known yet. Similarly, at least two isoforms of N. tabacum SAR1 proteins inhibits ER export in their GTP- or GDP-locked forms (Andreeva et al., 2000; daSilva et al., 2004; Osterrieder et al., 2010; Robinson et al.,

244

Yoko Ito et al.

2007; Schoberer et al., 2009; Yang et al., 2005). The expression of GDPlocked form of one of the N. tabacum SAR1s in BY-2 cells causes dissociation of SEC13 from the ER membrane (Yang et al., 2005). Recently, gene knockdown by RNAi of three rice SAR1 homologs specifically in the endosperm was shown to block ER export of seed storage proteins (Tian et al., 2013). The presence of COPII vesicles in plant cells is not fully established. Although there are a few reports finding vesicle-like structures by electron microscopy (Donohoe et al., 2007, 2013; Langhans et al., 2012; Staehelin and Kang, 2008), such structures are rarely found especially in highly vacuolated cells (Brandizzi and Barlowe, 2013; Hawes, 2012). This is thought to be due to rapid consumption of the vesicles at the ER–Golgi interface. It is impossible to rule out the possibility that ER–Golgi transport is mediated by direct connection by tubules; however, there are too few reports supporting this idea (Brandizzi et al., 2002; Hawes, 2012).

3.2.2 Cargo sorting During COPII vesicle formation, the prebudding complex interacts with cargo proteins and concentrates them in the budding vesicle (Kuehn et al., 1998). The role of Sec24 in cargo recognition has been revealed by structural, biochemical, and genetic approaches using yeast (Miller et al., 2003; Mossessova et al., 2003). As described above, there are three SEC24 isoforms in the Arabidopsis genome. Because Atsec24A null mutant is lethal, and the aberrant morphology of the ER and the Golgi caused by the point mutation in AtSEC24A cannot be suppressed by overexpression of AtSEC24B or AtSEC24C, these isoforms might be functionally different (Faso et al., 2009; Nakano et al., 2009). A further study showed that the loss of AtSEC24A function induces a defect in pollen germination but does not affect the female gametophyte (Conger et al., 2011). On the other hand, AtSEC24B and AtSEC24C were shown to function redundantly in both male and female gametogenesis (Tanaka et al., 2013). Moreover, microarray analysis of microspores indicates that the AtSEC24A expression peak comes earlier than that of AtSEC24B during pollen development (Conger et al., 2011). Since Arabidopsis SEC24 proteins are also thought to function in cargo sorting like the homologs in yeast, the difference between AtSEC24A and AtSEC24B/C may reflect specific cargo recognition. Indeed, such differentiation among Sec24 isoforms in cargo recognition has been reported for

Golgi Apparatus of Plant Cells

245

mammals and yeast (Kurihara et al., 2000; Merte et al., 2010; Pagano et al., 1999; Roberg et al., 1999; Wendeler et al., 2007). The particular recognition targets of AtSEC24 isoforms are yet to be characterized. 3.2.3 Uncoating, tethering, and fusion Unlike the COPI coat, Sec23, a subunit of the COPII coat complex, is known to act also as the GAP for Sar1 (Bi et al., 2002; Yoshihisa et al., 1993). In addition to the basic GAP activity of Sec23, Sec31 binding to both Sec23 and Sar1 accelerates Sar1-GTP hydrolysis by an order of magnitude (Antonny et al., 2001; Bi et al., 2007). Sar1 inactivation rapidly leads to COPII coat disassembly (Antonny et al., 2001). GTP hydrolysis is not required for COPII budding itself (Barlowe et al., 1994; Oka and Nakano, 1994) and was thought to be dispensable for cargo loading (Barlowe et al., 1994), but now it is known to contribute to efficient and proper cargo selection as described in Section 3.2.1. There are seven SEC23 isoforms in the Arabidopsis genome, greatly outnumber the only one Sec23 in yeast and two in human (Robinson et al., 2007). The localization of one of the AtSEC23s has been observed by immunofluorescence (Yang et al., 2005) and another by fluorescent protein tagging (see Section 3.2.4; Stefano et al., 2006), and the latter has been shown to bind AtSEC24A (Stefano et al., 2006). Both are believed to function similarly to the yeast homolog. The functional significance of AtSEC23 multiplicity is unknown. As the tethering factors for COPII vesicles, a long coiled-coil protein p115 (yeast Uso1) and a multisubunit complex transport protein particles I (TRAPPI) are well known. p115 and yeast homolog Uso1 have been revealed to be essential for COPII tethering. In the case of Uso1, it is thought to tether COPII vesicles directly to the cis-Golgi (Cao et al., 1998; Sapperstein et al., 1996). On the other hand, the target of COPII vesicle tethering by p115 is less clear in mammalian cells; in addition to the cis-Golgi cisternae, it can be other COPII vesicles and ERGIC (Allan et al., 2000; Szul and Sztul, 2011; Xu and Hay, 2004). p115 and Uso1 are recruited to COPII vesicles by the active form of the small GTPases Rab1 and its yeast homolog Ypt1, respectively (Allan et al., 2000; Cao et al., 1998). From yeast studies, TRAPPI complex is known to act as the GEF for Ypt1 (Wang et al., 2000). TRAPPI itself also interacts directly with Sec23 subunit through Bet3 subunit and function as a multisubunit tether (Cai et al., 2007; Sacher et al., 1998). In plants, homologs of p115 and all TRAPPI subunits have been found (Latijnhouwers et al., 2005). Arabidopsis p115 homolog was first isolated as

246

Yoko Ito et al.

GOLGIN CANDIDATE 6 (GC6), and the GFP-fusion protein or only the C-terminal domain tagged with GFP was shown to localize to the restricted domain of the Golgi stacks (Latijnhouwers et al., 2007). The localization of GC6/Atp115 has been also observed by immunoelectron microscopy, which locates GC6/Atp115 to the surface of COPII vesicles and the ribosome-exclusion area (Golgi matrix region) of the cis-Golgi cisternae in Arabidopsis root cells and tobacco BY-2 cells (Kang and Staehelin, 2008). Subsequently, the responsible gene for Arabidopsis maigo4 mutant, which accumulates seed storage proteins in the ER, was revealed to be GC6/Atp115 (Takahashi et al., 2010). The function of the Arabidopsis homolog of TRAPPI complex is yet to be investigated. The function of Ypt1/Rab1 is likely to be conserved also in plants, because Rab1 homologs from some plants have been reported to complement the yeast ypt1 mutant (Kim et al., 1996; Park et al., 1994). Rab1 homologs in higher plants are called RABD group, and divided into two subclasses, RABD1 and RABD2, according to their sequences (Pinheiro et al., 2009). Arabidopsis has one RABD1 and three RABD2 (RABD2a, RABD2b, and RABD2c), and the RABD2 members are thought to be extensively redundant. Both subclasses are involved in ER–Golgi trafficking because dominant negative forms of RABD1 or RABD2a inhibit ER export of secretory cargo and some Golgi or vacuolar proteins (Batoko et al., 2000; Park et al., 2004; Pinheiro et al., 2009; Saint-Jore et al., 2002; Samalova et al., 2006; Zheng et al., 2005). However, the dominant negative effects of a mutant protein of one subclass cannot be suppressed by overproduction of the wild-type protein of the other subclass, suggesting that these subclasses are functionally different. In support of this, RABD2 subclass is essential but RABD1 subclass is not. Nevertheless, rabD2b rabD2c double mutants are normal but rabD1 rabD2b rabD2c triple mutants are short and bushy with low fertility, indicating that these two subclasses have partially overlapping functions (Pinheiro et al., 2009). For membrane fusion of COPII vesicles to the cis-Golgi after tethering, the SNARE proteins Sed5 (Qa), Bos1 (Qb), Bet1 (Qc), and Sec22 (R-SNARE) are thought to be responsible in yeast (Barlowe and Miller, 2013). Nine Arabidopsis SNARE proteins have been localized to the Golgi apparatus by a systematic analysis using Arabidopsis cultured cells: Qa-SNARE AtSYP31 and AtSYP32; Qb-SNARE AtGOS11, AtGOS12, AtMEMB11, and AtMEMB12; Qc-SNARE AtBS14a and AtBS14b; and R-SNARE AtVAMP714 (Uemura et al., 2004). In addition, AtSEC22 expressed in tobacco leaf epidermal cells or protoplasts is shown to localize

Golgi Apparatus of Plant Cells

247

to the ER and the Golgi stacks (Bubeck et al., 2008; Chatre et al., 2005). AtSYP31 and AtSYP32 are Sed5 homologs, and AtMEMB11 and AtMEMB12 are homologs of membrin, the mammalian functional homolog of Bos1 (Sanderfoot et al., 2000). Overexpression of AtSYP31, MEMB11, or AtSEC22 has been reported to cause relocalization of Golgi proteins into the ER, which indicates their involvement in ER–Golgi transport (Bubeck et al., 2008; Chatre et al., 2005). AtBS14a and AtBS14b are similar to Bet1 and Sft1, another Golgi-localized Qc-SNARE from the same family of Bet1. The overexpression of either AtBS14a or AtBS14b is reported to support the growth of the yeast sft1 null mutant but not the bet1 mutant (Tai and Banfield, 2001). Therefore, AtBS14a and AtBS14b may be responsible for intra-Golgi trafficking like yeast Sft1 (see Section 4.2). 3.2.4 ERES COPII vesicles are known to bud from specialized domains of the ER membrane called transitional ER (tER) or ER exit/export sites (ERES). The existence of such domains with intermediate features between the rough ER and the Golgi vesicles was suggested long before the identification of the coated vesicles (Palade, 1975). It was later revealed that the COPII components are enriched in the ERES by immunoelectron microscopy in mammal and yeast cells, therefore, the COPII components became convenient markers for the ERES (Kuge et al., 1994; Orci et al., 1991; Paccaud et al., 1996; Rossanese et al., 1999). By immunofluorescence imaging or live imaging using fluorescent protein-tagged COPII components, the ERES appear as multiple discrete puncta (Hammond and Glick, 2000; Horton and Ehlers, 2003; Okamoto et al., 2012; Rossanese et al., 1999). Among the Sec proteins required for vesicle formation from the ER, the function of Sec16 has been long unknown. Sec16 localizes to the ERES and its inactivation causes disruption of the ERES in several species (Bhattacharyya and Glick, 2007; Connerly et al., 2005; Ivan et al., 2008; Watson et al., 2006; Yorimitsu and Sato, 2012). Sec16 is a large multidomain peripheral protein of the ER membrane, and binds to the COPII components through independent domains for each subunit (Espenshade et al., 1995; Gimeno et al., 1996; Shaywitz et al., 1997). From these facts, it has been speculated that Sec16 acts as the scaffold for COPII coat assembly at the ERES. Recently, Yorimitsu and Sato (2012) have reported that yeast Sec16 interferes with Sec31-mediated Sec23 GAP activation, which leads to stabilization of COPII coat complex and prevents premature complex

248

Yoko Ito et al.

disassembly. In addition, Sec16 alone has the ability to self-assemble into homooligomeric complexes (Yorimitsu and Sato, 2012). This Sec16 cluster is suggested to contribute as the basis of the ERES, and spatially restricts the vesicle budding. In the Arabidopsis genome, two Sec16 homologs are predicted (Robinson et al., 2007). Mammals also have two Sec16 homologs, one being about twice as long as the other. The two proteins are both necessary for normal ERES organization and ER export (Bhattacharyya and Glick, 2007). Two Arabidopsis SEC16 proteins are predicted to have almost the same size as the shorter counterpart of mammals (Robinson et al., 2007). A most recent study has revealed that one of them, SEC16A/MAG5, interacts with COPII coats and localizes to the ERES. Its mutant plants exhibit compromised ER export, indicating its role in ER-to-Golgi trafficking (Takagi et al., 2013). In the cells with discrete Golgi stacks such as P. pastoris, Caenorhabiditis elegans, and Drosophila melanogaster, the ERES and the Golgi stacks are almost always closely associated to form “secretory units” (Kondylis and Rabouille, 2009; Rossanese et al., 1999; Witte et al., 2011). In mammalian cells with the concentrated Golgi, the ERGIC is frequently associated close to the ERES instead of the Golgi complex (Appenzeller-Herzog and Hauri, 2006; Bannykh et al., 1996). Moreover, when the microtubules are disrupted by nocodazole treatment in mammalian cells, the Golgi ministacks are located near the ERES like in invertebrate cells (Hammond and Glick, 2000). Also in plant cells, many punctate signals can be observed by using COPII coat subunits as markers. However, conflicting results have been reported about the relationship between the ERES and the Golgi stacks. By immunofluorescence observation of COPII coat proteins, the signals labeled many punctate structures greatly outnumbering the Golgi stacks. Some of them were associated with the Golgi stacks, and such signals were often found on the areas of the ER tubules associated with the Golgi stacks. Furthermore, by live imaging of tomato SEC13 tagged with GFP in tobacco BY-2 cells, the signal pattern was similar to the immunofluorescence, and the Golgi stacks were observed to associate with the SEC13-GFP signal around their rims. The level of SEC13-GFP around the Golgi stacks varied along the time, and the Golgi stacks moving slower tended to have more SEC13 signals (Yang et al., 2005). GFP-tagged tobacco SAR1, which can inhibit ER–Golgi transport in the dominant negative mutant form, also labeled puncta either associated or nonassociated with the Golgi stacks

Golgi Apparatus of Plant Cells

249

(Robinson et al., 2007). From these data, the “kiss-and-run” model was suggested: the ERES are relatively stable, and the Golgi stacks temporarily associates with them and ER-to-Golgi transport takes place during their contacting periods (Fig. 6.4A). On the other hand, tobacco SAR1 and Arabidopsis SAR1, SEC23, SEC24, and SEC13 fused with fluorescent proteins marked puncta associated with the Golgi stacks, without signals independent from the Golgi. The COPII signals moved always with the associated Golgi stacks (daSilva et al., 2004; Hanton et al., 2007, 2008, 2009; Stefano et al., 2006). The punctate signals of YFP-SEC24 increased in number and became more definite by expression of the cargo with COPII-dependent ER export motif, indicating the involvement of the puncta of YFPSEC24 in ER export (Hanton et al., 2007). Moreover, FRAP analysis showed that ER-to-Golgi protein transport can occur while the Golgi stacks are moving (daSilva et al., 2004). Therefore, the “secretory unit” model was suggested: the Golgi stacks and the ERES are continuously associated and the ER-to-Golgi transport continues during rapid movement (Fig. 6.4B). Our recent study has demonstrated that Arabidopsis SEC13 fused with YFP expressed in tobacco BY-2 cells labeled many puncta outnumbering the Golgi stacks, and some of them are associated with the Golgi in ring-like shapes surrounding the Golgi rims. Moreover, the SEC13-YFP signal moved together with the associated Golgi stacks (Ito et al., 2012). These results suggest that the truth is in between the two extreme models (hybrid model, Fig. 6.4C). Because the SEC13-YFP puncta without associated Golgi stacks were smaller and extremely mobile independently of actin filaments, we speculate that they become stable and active when they encounter the Golgi stacks. Recent work suggested that the COPII signal is not on the ER but in the small gap between the ER and the Golgi, which represents the COPII vesicles accumulating at the cis-face of the Golgi, not the ERES (Langhans et al., 2012). In Arabidopsis leaf epidermal cells, pulling of individual Golgi stacks by a focused infra-red laser beam (laser tweezers) was shown to drag the ER membrane as tubules behind them (Sparkes et al., 2009b). This result indicates the physical connection between the Golgi and the ER, which involves Golgi matrix and tethering proteins. Such matrix proteins are thought to catch the COPII vesicles before fusion with the cis-Golgi while the Golgi stacks are moving, therefore Langhans et al. (2012) suggested a difficulty to know whether the ERES move together with the Golgi stacks or not by the observation of only COPII coats. Furthermore, COPI vesicles responsible for Golgi-to-ER transport are proposed to fuse with the ER

250

Yoko Ito et al.

Figure 6.4 Models for ERES–Golgi organization in plants cells. (A) The kiss-and-run model. In this model, the ERES are relatively stable, and the Golgi stacks travel from one ERES to another. ER-to-Golgi transport takes place during the temporal association of the Golgi stacks and the ERES. (B) The secretory unit model. In this model, the Golgi stacks and the ERES are continuously associated and move together. ER-to-Golgi transport occurs not only during their pauses but also during rapid movement. (C) The hybrid model. In this model, some of the ERES are continuously associated with the Golgi stacks similarly to the secretory unit model, but the others are not. The ERES without associated Golgi stacks are smaller and move independently from the Golgi stacks. They get stable and active when they encounter the Golgi stacks. (D) The modified secretory unit model. In this model, the temporal pauses of the Golgi stacks take place at the ERIS. The COPII and COPI vesicles are continuously formed during the Golgi movement, but their fusion with the target membranes only occurs while the Golgi stacks are immobile. Both types of vesicles accumulate between the Golgi and the ER, and move together with the Golgi stacks.

Golgi Apparatus of Plant Cells

251

during temporary pauses at “ER import sites” (ERIS, see Section 3.1.5; Lerich et al., 2012). These data are summarized and called as the modified secretory unit model (Fig. 6.4D; Lerich et al., 2012). Unlike the COPII coat subunits, the distribution of Sec12 on the ER membrane is different between species. Sec12 is concentrated to the ERES near the Golgi stacks in P. pastoris, whereas Sec12 is dispersed throughout the ER in S. cerevisiae (Rossanese et al., 1999). Because AtSEC12-YFP expressed in tobacco cells localizes throughout the ER, it is not appropriate as an ERES marker in plant cells (daSilva et al., 2004). Analysis using SEC16 should shed light on the controversy. The plant Golgi stacks are observed associated with the ER tubules and the rims of the ER cisternae, but not with the ER cisternal surface (Fig. 6.5; Sparkes et al., 2009a). If active ERES continuously associate and move with the Golgi stacks, the ERES should locate to the ER domains with high membrane curvature. Such curvature of the ER is at least partially regulated by reticulon family proteins (Sparkes et al., 2010). Recently, Okamoto et al. (2012) have revealed that the high curvature of the ER membrane, especially saddle-like structures that contain both positive and negative curvatures, is important for ERES localization in yeast. In the S. cerevisiae cells lacking reticulon family proteins, peripheral ER network was disrupted, and the ERES were localized to the remaining few high-curvature domains. Moreover, the dynamics and morphology of the Golgi were affected in the mutants. Similar regulation may well exist in plant cells.

Figure 6.5 Association of the ER tubules and the Golgi stacks. Confocal image of a tobacco BY-2 cell expressing SP-GFP-HDEL (ER lumen, green) and mRFP-AtSYP31 (cis-Golgi, magenta). The ER tubules seem to surround the Golgi stacks on the cisternal rims. Scale bar, 10 mm.

252

Yoko Ito et al.

3.3. ERGIC Compared to plants, fungi, and invertebrates, the Golgi apparatus is far from the ERES in mammalian cells because the Golgi is concentrated near the nucleus. ERGIC consists of vesiculotubular clusters, and is marked by a lectin ERGIC-53 (Appenzeller-Herzog and Hauri, 2006; Schweizer et al., 1990). Observations of the glycoprotein cargo transport revealed that the structure is the intermediate compartment of transport between the ER and the Golgi (Saraste and Svensson, 1991; Schweizer et al., 1990). In the ERGIC, sorting and concentration of anterograde and retrograde cargo by COPI vesicles are thought to take place (Martı´nez-Mena´rguez et al., 1999). About the transport between the ER and the Golgi via the ERGIC, there are two models. In the “transport complex model,” ER-derived COPII vesicles fuse with each other to form the ERGIC, and the ERGIC travels to the Golgi on the microtubule to become the cis-Golgi by fusing with other ERGIC or to fuse with preexisting cis-Golgi. This model is supported by the observations that ERGIC moves from the ERES to the Golgi along microtubules (Presley et al., 1997; Scales et al., 1997). On the other hand, in the “stable compartment model,” ERGIC is a static structure and facilitates a two-step transport process. First, COPII vesicles mediate short-range transport from the ER to the ERGIC independently of microtubules, and next the anterograde cargo-rich domains of the ERGIC are transported long distance to the cis-Golgi depending on microtubule tracks. A live imaging study revealed that the ERES was stationary and small soluble cargo but not ERGIC-53 was exported to the Golgi (Ben-Tekaya et al., 2005). It is also possible that different mechanisms apply for different cargo proteins. In the cells whose Golgi stacks are closely associated to the ERES, the presence of ERGIC has not been presumed. However, a recent report suggests that C. elegans cells have ERGIC-like compartments between the cisGolgi and the ERES (Witte et al., 2011). Also in plant cells, it is proposed that the cis-most Golgi cisternae are biosynthetically inactive and function as the site of membrane assembly and cargo sorting, similarly to the mammalian ERGIC (Donohoe et al., 2013). Moreover, we have demonstrated by BFA treatment and its removal of tobacco BY-2 cells that punctate structures with some cis-Golgi components near the ERES act as the scaffold for Golgi stack regeneration, suggesting their ERGIC-like properties (see Section 5.4; Ito et al., 2012).

Golgi Apparatus of Plant Cells

253

4. INTRA-GOLGI TRAFFICKING 4.1. Models of intra-Golgi trafficking To discuss the formation of the Golgi structure, we cannot skip the long-lasting debate about intra-Golgi trafficking (Glick and Luini, 2011; Glick and Nakano, 2009; Nakano and Luini, 2010). This debate is about how cargo molecules are transported through the stack of cisternae, and has involved many cell biologists. Among many models proposed, two major models were the issue of dispute; one is the vesicular transport (stable compartments) model, and the other is the cisternal maturation model (Fig. 6.6).

Figure 6.6 Two major models for intra-Golgi trafficking. (Left) The vesicular transport (stable compartments) model. In this model, the Golgi cisternae are stable and distinct suborganelles with different resident proteins. Cargo proteins are transported from one cisterna to the next by anterograde COPI vesicles, whereas the Golgi proteins are excluded from them and remain in the cisternae. Cargo proteins reach the TGN and exit by clathrin-coated vesicles or other carriers. (Right) The cisternal maturation model. In this model, the Golgi cisternae themselves function as the cargo carriers. They are transient compartments that are newly formed by homotypic fusion of COPII vesicles, progress from cis to trans, and break down at the TGN stage. The Golgi proteins recycle back from later to earlier cisternae by retrograde COPI vesicles, and the nature of the cisternae gradually changes as they progress.

254

Yoko Ito et al.

In the vesicular transport model, the Golgi stack is viewed as a series of distinct suborganelles, each of which has a characteristic set of resident proteins. Cargo proteins travel from one cisterna to the next by anterograde vesicles, and the Golgi resident proteins remain in the cisterna by exclusion from the vesicles. The fundamental idea of this model can be traced back to the data which demonstrated that the peri-Golgi vesicles contained newly synthesized secretory proteins ( Jamieson and Palade, 1966). As described in Section 3.1.1, cell-free reconstitution assays and cell fusion experiments in 1980s indicated vesicular transport between cisternae, and subsequently identified COPI vesicles were thought to be responsible for this transport. The model was updated later in which COPI vesicles move bidirectionally, with anterograde vesicles carrying secretory cargo and retrograde vesicles recycle trafficking components (Orci et al., 2000b). On the other hand, the cisternal progression model, which was upgraded later to the cisternal maturation model, was first proposed from morphological studies (Grasse´, 1957). In this model, Golgi cisternae themselves are thought as the cargo carriers. Homotypic fusion of ER-derived COPII vesicles or other carriers like ERGIC nucleates the new cis-cisternae, which progress forward along the secretory pathway to become the TGN. The Golgi cisternae are transient compartments and continuously turn over. In the updated cisternal maturation model, COPI vesicles recycle Golgi resident proteins from later to earlier cisternae, and the nature of the cisternae gradually changes as they progress (maturation). This model was supported by the fact that the cargoes much larger than COPI vesicles can travel across the Golgi stacks. Cell surface scales of algae are assembled in the Golgi apparatus, and transported forward without being packed into COPI vesicles (Becker et al., 1995; Brown et al., 1970; Melkonian et al., 1991). Procollagen aggregates in mammalian fibroblasts are also shown to progress across the Golgi stacks without getting into small vesicles (Bonfanti et al., 1998; Mironov et al., 2001). One important problem here is the direction of intra-Golgi COPI vesicles. Many studies have been made using mammalian cells to solve this problem, but have led to conflicting conclusions. Some found that Golgi glycosylation enzymes were enriched in COPI vesicles but cargo proteins were not (Gilchrist et al., 2006; Martı´nez-Mena´rguez et al., 2001), whereas others reported the opposite (Cosson et al., 2002; Kweon et al., 2004; Orci et al., 2000a). Recent studies on the tethering factors involved in intra-Golgi movement of COPI vesicles revealed that

Golgi Apparatus of Plant Cells

255

untethered COPI vesicles with Golgi resident proteins accumulated when they were knocked down (Sohda et al., 2010; Zolov and Lupashin, 2005). Only retrograde but not anterograde trafficking was disrupted in these experiments, indicating the involvement of COPI vesicles in the intra-Golgi retrograde trafficking. It was also suggested that there are at least two types of COPI vesicles according to their tethering factors, and this differentiation can be coupled with their transport directions (forward/bidirectional and backward) (Malsam et al., 2005; Orci et al., 2000b). However, the two COPI populations could be retrograde carriers, and the existence of anterograde vesicles remains to be confirmed. Recent data from S. cerevisiae by high-resolution confocal microscopy have revealed that the protein components change in a single cisterna from cis-Golgi proteins to trans-Golgi proteins over time, which gives a strong support for the maturation of cisternae (Losev et al., 2006; MatsuuraTokita et al., 2006). Detailed morphological studies by electron tomography of the cells of human, Pichia pastoris, Arabidopsis thaliana, Scherffelia dubia, Chlamidomonas reinhardtii, and Dionaea muscipula (Venus flytrap) support cisternal maturation (Donohoe et al., 2013; Mogelsvang et al., 2003; Storrie et al., 2012). Moreover, observations of Golgi regeneration after BFA treatment and removal in mammalian and plant cells have revealed that the Golgi stacks regenerate in the cis-to-trans order, which is consistent with the cisternal maturation model (see Section 5.4; Alcalde et al., 1992; Ito et al., 2012; Puri and Linstedt, 2003; Schoberer et al., 2010). Therefore, the basic concept of the cisternal maturation model appears to be widely accepted now. However, the debate has not come to an end yet. Patterson et al. (2008) reported that both big and small secretory cargoes exit the Golgi with exponential kinetics in mammalian cells. This appeared to conflict with the cisternal maturation, in which the cargoes are predicted to exit the Golgi with linear kinetics. The problem can be avoided by modifying the model to include a post-Golgi compartment such as TGN as longlived organelle, but so far there is no supporting evidence. Moreover, recent two studies using similar procedures report completely opposite results (Lavieu et al., 2013; Rizzo et al., 2013). The two groups used the domain which can be polymerized at will, and one group fused it with the CTS region of Golgi mannosidase and the other fused it with the luminal side of a plasma membrane protein. The polymerized protein of the former group progressed forward across the Golgi stack,

256

Yoko Ito et al.

whereas that of the latter group stayed and was not transported to later cisternae. The debate is still going on, and we obviously need further investigations.

4.2. Tethering and fusion of intra-Golgi COPI vesicles For tethering of intra-Golgi COPI vesicles, although the precise origin and transport direction of the vesicles are not clear, the coiled-coil tethers p115 (yeast Uso1), giantin, golgin-84, and CASP (yeast Coy1), and the multisubunit complexes TRAPPII and COG are suggested to be involved. Uso1 is essential only for COPII tethering as described in Section 3.2.3, but mammalian p115 additionally binds to b-COP and is involved in COPI tethering (Guo et al., 2008). Giantin and golgin-84 bind to p115 and CASP, respectively, on the Golgi membranes (Malsam et al., 2005). COG is an octameric complex composed of Cog1–8 subunits, and yeast Cog2 and human Cog3 have been reported to interact with g- and b-COP, respectively (Suvorova et al., 2002; Sztul and Lupashin, 2009; Vasile et al., 2006; Zolov and Lupashin, 2005). Yeast COG complex strongly interacts with active Ypt1, and also with Ypt6 to a less extent independently of nucleotides (Suvorova et al., 2002). Similarly, mammalian COG complex interacts with Rab1 and Rab6, the homologs of Ypt1 and Ypt6 (Fukuda et al., 2008). Not only p115 and COG complex but also giantin and golgin-84 interact with active Rab1 (or yeast Ypt1; Beard et al., 2005; Diao et al., 2003; Fukuda et al., 2008; Satoh et al., 2003). Mammalian TRAPPII is suggested to act not only as a tether but also as a specific GEF of Rab1 similarly to TRAPPI, and the activated Rab1 recruits effectors including the tethering factors above (Allan et al., 2000). TRAPPI and TRAPPII complexes share common subunits, and TRAPPII has additional three subunits, which are thought to determine the binding specificity different from TRAPPI (Sztul and Lupashin, 2009). Mammalian TRAPPII complex is especially associated with COPI vesicles and buds, and its subunits interact with g-COP (Yamasaki et al., 2009). In plants, homologs of p115, golgin-84, CASP, and COG complex components have been found, but not for giantin and the TRAPPII subunits that are not shared with TRAPPI (Latijnhouwers et al., 2005). Although the role of Atp115 in ER-to-Golgi anterograde transport is

Golgi Apparatus of Plant Cells

257

indicated as described in Section 3.2.3, it is unknown whether it mediates COPI tethering. Homologs of golgin-84 were identified as GC1 and GC2, and their GFP-fusion proteins localized to the cis-Golgi cisternal rims (Latijnhouwers et al., 2007). The Arabidopsis CASP homolog AtCASP was identified based on similarity to human CASP in transmembrane region, and localized to the Golgi cisternal rims in tobacco leaf epidermal cells (Latijnhouwers et al., 2007; Renna et al., 2005). This localization was later revealed to be the cisternae between those with the cis-Golgi SNARE AtSYP31 and the trans-Golgi marker ST in tobacco BY-2 cells (Ito et al., 2012). The physiological roles of GC1, GC2, and AtCASP in plant cells have not been reported. For COG complex, homologs of all subunits are coded in the Arabidopsis genome (Latijnhouwers et al., 2005). Among the eight subunits, the gene coding Cog7 has been shown to be responsible for embryo yellow (eye) mutant, which exhibits defects in cell expansion and organization of the shoot apical meristem. In eye mutant embryos, ERD2-GFP mislocalizes to the ER and GFP-tagged cis-Golgi localized SNARE MEMB12 shows aberrant shaped punctate structures (Ishikawa et al., 2008). In addition to Rab1 homolog RABDs (see Section 3.2.3), Rab6/Ypt6 group is also conserved in plants; the Arabidopsis homologs are called RABH group (Saito and Ueda, 2009). Because one of the RABH proteins AtRABH1b has been reported to replace the function of yeast Ypt6, the molecular function is thought to be conserved (Bednarek et al., 1994). After tethering, membrane fusion is thought to be mediated by SNARE proteins; Qa-SNARE Sed5 (mammalian syntaxin 5), Qb-SNARE Gos1 (Gos28), Qc-SNARE Sft1 (GS15), and R-SNARE Ykt6 or Sec22 (Barlowe and Miller, 2013; Kloepper et al., 2007; Shorter et al., 2002; Xu et al., 2002). Among the plant Golgi-localized SNAREs (see Section 3.2.3), AtSYP3, AtGOS1, and AtBS14 families, and AtSEC22 are presumed to be involved in intra-Golgi COPI tethering from their homology to yeast SNAREs (Sanderfoot et al., 2000; Tai and Banfield, 2001). Because overexpression of AtBS14a or At BS14b has been shown to partially suppress the phenotypes of yeast sft1 deletion mutant, their functions are thought to be conserved to some extent (Tai and Banfield, 2001). For AtSYP31 and AtSEC22, although their involvement in ER–Golgi trafficking is speculated, there is no experimental data indicating its role in intraGolgi transport.

258

Yoko Ito et al.

5. FORMATION AND MAINTENANCE OF STACKED STRUCTURE 5.1. Golgi apparatus as stacked cisternae As described in Section 2.1, the unique stacked structure of the Golgi apparatus is conserved among nearly all eukaryotes. Certainly, there are a few exceptions such as S. cerevisiae, which has unstacked Golgi, and some unicellular parasites, which appear to lack Golgi. Many of such “Golgi-lacking” organisms also lack identifiable mitochondria and peroxisomes, and they were thought to have evolved very early before the innovation of the Golgi in the early studies (Cavalier-Smith, 1987). However, improved phylogenetic analyses revealed later that all of them possess most genes encoding the proteins associated with trafficking around the Golgi such as coat proteins, Rab GTPases, and SNAREs. Moreover, the organisms without stacked Golgi like S. cerevisiae are spread throughout the eukaryotic tree, each of which is located in the clade of organisms with the Golgi stacks (Klute et al., 2011; Mowbrey and Dacks, 2009). Thus, the loss of the normal Golgi structures is now thought to be secondary events. The phylogenetic tree suggests that Golgi stacking has been lost at least eight times independently (Mowbrey and Dacks, 2009). Thus, the ancestral eukaryote must have had stacked Golgi and the structure has been inherited to most descendants during evolution. This indicates that the stacked structure is important for the Golgi function and therefore for better survival of eukaryotes. The relationship between the Golgi structure and functions has been asked for a long time. Cisternal stacking may contribute to efficient transport through the Golgi apparatus, because the donor and target membranes are physically close to each other. However, S. cerevisiae without stacked Golgi achieves efficient protein secretion (Nakano and Luini, 2010). Also in Drosophila, the cells at many developmental stages are known to exhibit no Golgi stacks but instead possess vesiculotubular clusters, still achieving efficient secretion (Kondylis and Rabouille, 2009). By contrast, Golgi stacking might limit the intra-Golgi transport. COPI vesicles are known to form exclusively at the cisternal rims, and different behaviors between the rims and the center of the cisternae are proposed (Lavieu et al., 2013). Golgi unstacking can increase the membrane surface of cisternae, which facilitates interaction with coat proteins to produce transport vesicles. In support of this, artificially unstacked mammalian

Golgi Apparatus of Plant Cells

259

Golgi was reported to produce vesicles more efficiently than normal Golgi stacks in vitro (Wang et al., 2008). From another viewpoint, Golgi stacking is presumed to ensure complex cargo glycosylation by sequential exposure of cargo to enzymes. Some “Golgi-lacking” organisms are suggested to possess fewer ER glycosyltransferases compared to their relatives with Golgi stacks, implying their difference in Golgi glycosylation as well (Samuelson et al., 2005). A recent study shows that disruption of the Golgi structure accelerates trafficking but impairs proper cargo glycosylation (Xiang et al., 2013). Further comparative studies will give us more insights.

5.2. Search for the “Golgi stacking factors” In order to form and maintain the stacked structure of the Golgi apparatus, there should be molecular factors responsible for it. Such components have been sought for a long time. By electron microscopy, bridge-like components cross-linking the adjacent cisternae were observed in animal and plant cells, which turned out to be proteins sensitive to proteases (Cluett and Brown, 1992; Franke et al., 1972; Mollenhauer, 1965; Staehelin et al., 1990). To identify these components, the procedures used for characterization of nuclear lamina were applied; membranes of isolated Golgi stacks were solubilized by detergent, and after removal of Golgi enzymes so-called Golgi matrix was isolated (Slusarewicz et al., 1994). Antisera raised against this matrix led to identification of GM130 (Nakamura et al., 1995). An in vitro assay isolated another component. Mammalian Golgi apparatus disassembles upon mitosis and reassembles in the daughter cells at telophase, and this Golgi disassembly and reassembly can be mimicked in a cell-free assay treating the isolated Golgi apparatus with mitotic or interphase cytosol. An alkylating agent NEM (N-ethylmaleimide) prevented Golgi reassembly, and its major target turned out to be a 65-kDa protein. The protein was named GRASP65 (Golgi ReAssembly Stacking Protein), because inhibiting its function by addition of its antibody or recombinant GRASP65 after Golgi disassembly led to the formation of single cisternae without stacking upon reassembly (Barr et al., 1997). GRASP65 was revealed to be enriched in the detergent-resistant matrix fraction and interact with GM130 (Barr et al., 1998). Identification of GRASP65 led to the discovery of a homolog GRASP55 with a similar role (Shorter et al., 1999). The in vivo function of GRASP65 in Golgi stacking was supported to some extent by the experiments of microinjection of the antibody against

260

Yoko Ito et al.

GRASP65 and gene knockdown by RNAi (Su¨tterlin et al., 2005; Wang et al., 2003). However, upon inhibition of either GRASP65 or GRASP55 expression by siRNA, the Golgi ribbons fragmented into ministacks but Golgi stacking remained unaffected (Feinstein and Linstedt, 2008; Puthenveedu et al., 2006). Subsequently, double depletion of GRASP65 and GRASP55 by siRNA was performed, and electron microscopy revealed that stacked Golgi was absent in over 80% of the cells (Xiang and Wang, 2010). From these results, mammalian GRASPs are now thought to act in Golgi stacking redundantly, but independently in Golgi ribbon formation. In both cases, GRASPs probably function as bridges between adjascent compartments by oligomerization (Lowe, 2011). GM130 is also suggested to be required for both Golgi stacking and Golgi ribbon formation, presumably as the tethering factor acting before connection by GRASPs (Marra et al., 2007; Puthenveedu et al., 2006; Shorter and Warren, 1999). In spite of their prominent roles in the Golgi structure formation in mammalian cells, the function of GRASP homologs as the “Golgi stacking protein” is not very obvious in other organisms. Unlike vertebrates with clear Golgi ribbons, most organisms with discrete Golgi stacks (or no stacks) possess only one GRASP homolog (Vinke et al., 2011). In Drosophila cells, depletion of the only GRASP homolog dGRASP effects only 30% of the Golgi stacks, whereas double depletion with GM130 homolog increases the effect (Kondylis et al., 2005). In addition, dGRASP localizes to the Golgi membranes even when the Golgi cisternae are unstacked (Kondylis and Rabouille, 2009). Deletion of the GRASP homolog Grh1 in P. pastoris has no obvious effect on the Golgi structure (Levi et al., 2010). S. cerevisiae also has a GRASP homolog, which localizes to the cis-Golgi membrane, in spite of the well-known unstacked Golgi (Behnia et al., 2007). Surprisingly, GRASP is lacking in plants. Green algae also lack GRASP but red algae have it, indicating the secondary loss of GRASP in the chloroplastida stem lineage (Klute et al., 2011). Similarly, GM130 has not been found in plants either (Latijnhouwers et al., 2007). Needless to say, however, plants have stacked Golgi apparatus. Whether conservation of the genes involved is too low or completely different components substitute GRASP and GM130 in plants remain obscure. The plant Golgi stacks keep intact structure during their rapid movement and never break down while traveling (Madison and Nebenfu¨hr, 2011). The best candidate that keeps the structural integrity is the Golgi matrix proteins. By electron microscopy, plant cells show a zone surrounding the Golgi

Golgi Apparatus of Plant Cells

261

stack and the TGN, from which ribosomes are excluded (Staehelin and Kang, 2008; Staehelin et al., 1990). This “ribosome-exclusion zone” is thought to represent the Golgi matrix holding the Golgi structure together. Some matrix components have been identified in plant cells, including AtCASP and Atp115, which also act as vesicle tethering factors (Sections 3.2.3 and 4.2), but their functions in the maintenance of the Golgi structure are not known. Further investigations would reveal the mechanism for Golgi structure maintenance in plant cells, perhaps not much conserved with animals and yeast.

5.3. Golgi biogenesis upon mitosis Upon cell division, not only DNA but also organelles must be inherited to the daughter cells. The Golgi apparatus is not an exception. Some parasitic protozoa have only one Golgi stack per cell, which makes it easy to track the behaviors of the single Golgi. By several studies, the Golgi apparatus in such organisms are shown to double before cytokinesis by distinct procedures. In Trichomonas, the Golgi stack elongates laterally during cell cycle and undergoes medial fission before mitosis. The fission appears to occur from trans-to-cis direction in electron microscopy (Benchimol et al., 2001). In Toxoplasma gondii, similar lateral growth and fission are followed by an additional fission. The resulting four stacks are partitioned to the daughter cells in pairs, which subsequently coalesce to form a single Golgi in each daughter cell (Pelletier et al., 2002). Differently from these species, the new Golgi stack is formed de novo near the old Golgi in Trypanosoma brucei. The Golgi formation is coupled with duplication of the ERES, and the new Golgi stack is supplied with materials not only by the new ERES but also by the old Golgi (He et al., 2004). The new Golgi formation was visualized by using several Golgi proteins, showing that the Golgi matrix and enzymes are laid down first, followed by the association of COPI components and initiation of cargo trafficking (Ho et al., 2006). Such de novo formation of the ERES– Golgi unit has been also observed in P. pastoris, whereas the Golgi formation is not limited around mitosis. This process is seemingly independent of preexisting Golgi stacks, but it is also possible that there are invisible Golgi templates (Bevis et al., 2002; Rossanese et al., 1999). In contrast to the protozoans, the mammalian Golgi apparatus behaves in a completely different way. The single Golgi ribbon is initially disassembled into the ministacks in the late G2 phase, and even the stacks disappear at the early phase of mitosis. What occurs in the latter step is still under debate;

262

Yoko Ito et al.

whether the Golgi cisternae fragment into small vesicles, or the Golgi components are absorbed into the ER (Axelsson and Warren, 2004; Persico et al., 2009; Zaal et al., 1999). In either case, the Golgi components are dispersed throughout the cell and partitioned into the daughter cells, followed by reassembly of the Golgi ribbon in the telophase. The reassembly process has been intensively studied to reveal the mechanism of Golgi biogenesis, and led to the identification of GRASPs as described in the previous section. In plant cells, it is known that the Golgi stacks remain intact all through the cell cycle (Dupree and Sherrier, 1998; Nebenfu¨hr et al., 2000). Whether the Golgi apparatus disassembles or remains intact during mitosis is not species dependent but cell-type dependent, because Drosophila Golgi undergoes disassembly and reassembly in tissue cultured cells but keeps intact in embryonic cells (Stanley et al., 1997). It is hypothesized that the difference might depend on the ER–Golgi transport activity. During mitosis, Sec13 dissociates from the ERES and the ER export is arrested in mammalian cells (Hammond and Glick, 2000). On the other hand, plant cells require active secretion for cytokinesis, which is achieved by the cell plate. Cell plate formation needs new membrane and wall materials supplied by Golgi-derived vesicles to the phragmoplast, therefore disruption of the Golgi apparatus by BFA during mitosis results in binuclear cells (Reichardt et al., 2007). Because the plant Golgi stacks do not disassemble at any stages, they must somehow increase in number during the cell cycle. However, the process of Golgi biogenesis and its relationship to mitosis are almost completely unknown in plant cells. In onion root meristem cells, there is a report that the Golgi number increases mainly between prophase and metaphase (Garcia-Herdugo et al., 1988). On the other hand, the Golgi stacks in Catharanthus roseus cells laterally enlarge during cytokinses, and some stacks seemed to divide by electron microscopy (Hirose and Komamine, 1989). By electron tomography, the “dividing” Golgi looks as two stacks sharing a big trans-cisterna (Staehelin and Kang, 2008). According to the cisternal maturation model, such division should start with assembly of two half-sized cis-cisternae on the cis-surface of single enlarged stack, and maturation of the cis-cisternae would progress without fusion being followed by assembly of new cisternae, each of which eventually results in two separated stacks. However, electron microscopy cannot distinguish whether the stacks are dividing or fusing. A careful stereological study by electron microscopy in Arabidopsis shoot apical meristem cells has suggested that the Golgi stacks doubles in number during the G2 phase prior to mitosis, but its mechanism

Golgi Apparatus of Plant Cells

263

remains unknown (Seguı´-Simarro and Staehelin, 2006). Live cell imaging analysis also has been performed, however, no dramatic increase of the number of the Golgi has been observed in any specific time period (Nebenfu¨hr et al., 2000).

5.4. Golgi regeneration after BFA treatment As described above, BFA treatment causes disassembly of the Golgi apparatus and absorbance of most Golgi components into the ER in many types of cells, and this effect is reversible; the Golgi apparatus reassembles after BFA removal. This reassembly process has been investigated to reveal the mechanism of Golgi biogenesis. Especially in mammalian cells, many studies using BFA have been performed to know whether the Golgi biogenesis occurs de novo from the ER or requires preexisting templates. The candidate of such templates is the Golgi matrix, because it is segregated from other Golgi components including Golgi enzymes by BFA treatment, localizing to the “Golgi remnants” distinctive from the ER. In addition, mitosis progresses normally under this condition, and the Golgi apparatus is regenerated by BFA removal after cell division (Seemann et al., 2002). However, combination of BFA and another drug H89, which blocks ER export, induces ER redistribution of both matrix and nonmatrix proteins, and the Golgi reassembles normally by removing them (Puri and Linstedt, 2003). Therefore, the mammalian Golgi is thought to have the ability to form de novo under a certain condition. In plant cells, the behavior of the Golgi proteins upon BFA treatment has been controversial. In tobacco leaf epidermal cells, the Golgi proteins were reported to redistribute to the ER in the trans-to-cis order by BFA treatment. In addition, one of the matrix protein, AtCASP, did not fully relocate to the ER but localized to small punctate structures colocalizing with ERES (Schoberer et al., 2010). On the other hand, we and other have shown that both cis- and trans-Golgi cisternae collapse almost at the same time in tobacco BY-2 cultured cells (Ito et al., 2012; Madison and Nebenfu¨hr, 2011). Moreover, we have demonstrated that AtCASP relocalizes to the ER almost completely by BFA treatment (Ito et al., 2012). We have also demonstrated that the cis-Golgi proteins AtSYP31 and AtRER1B relocalize to punctate structures by BFA treatment in BY-2 cells, whereas another cis-Golgi protein AtERD2 is absorbed into the ER (Ito et al., 2012). Very similar results were obtained in Arabidopsis cultured cells by transient expression of dominant mutants of ARF1 (Takeuchi et al., 2002).

264

Yoko Ito et al.

The punctate structures of AtSYP31 and AtRER1B are smaller in size and larger in number compared to normal Golgi stacks, and detailed 3D analysis by super-resolution confocal live imaging microscopy (SCLIM) that we had developed has revealed that these structures lie very close to the ERES but are not identical to the ERES (Ito et al., 2012). Relation of these punctate structures with those of AtCASP observed in tobacco leaf epidermal cells (Schoberer et al., 2010) is unknown. It is also reported that expression of the GTP-locked dominant mutant of SAR1 in tobacco cells results in Golgi fragmentation and accumulation of vesiclulotubular clusters as the remnants (Osterrieder et al., 2010). Such clusters might be related to the punctate structures we observe, but further investigations are needed. After BFA removal, the punctate structures of the cis-Golgi proteins gather first, and subsequently the medial- and trans-Golgi markers follow sequentially to form Golgi stacks (Ito et al., 2012). This result indicates that the punctate structure function as “seeds” of Golgi regeneration by receiving cis-Golgi materials to form the cis-cisternae, and regeneration progresses in the cis-to-trans order (Fig. 6.7). According to a report by electron tomography, the cis-most cisternae in plant cells are biosynthetically inactive and might function similarly to mammalian ERGIC (Donohoe et al., 2013). In mammalian cells upon BFA treatment, not only matrix proteins but also some cis-Golgi membrane proteins that cycle between the ER and the Golgi and also the ERGIC protein ERGIC53 localize to the Golgi remnants (Fu¨llekrug et al., 1997a,b; Lippincott-Schwartz et al., 1990; Tang et al., 1993). From these facts, we speculate that the cis-most cisternae in plant cells function like both the ERGIC and the cis-Golgi of mammalian cells, and the proteins localizing there relocate to the punctate structures by BFA treatment that act as the scaffold for Golgi regeneration. In the later stage of Golgi regeneration after BFA removal, the Golgi stacks keep growing until they reach the size twice as large as normal Golgi, and subsequently divide by fission (Hawes et al., 2008a; Langhans et al., 2007). By electron microscopy, the fission seems to occur in the cis-to-trans order, consistent with the fission by maturation (Section 5.3). It is still unknown in which way the new Golgi is formed (duplication by fission, de novo formation from the ER, or the intermediate), but these results from the Golgi regeneration experiments indicate that any of them can occur in plant cells, although the final division has not been observed in living cells. Further studies by live imaging with higher time and space resolution will provide us more insights.

Golgi Apparatus of Plant Cells

265

Figure 6.7 A model for the behaviors of the tobacco Golgi apparatus upon BFA treatment and removal. (A) The Golgi stacks are located near the ERES, and there might be some ERES without associated Golgi. (B) By BFA treatment, some cis-Golgi proteins localize to the punctate compartments closely associated to the ERES, while other cis-, medial-, and trans-Golgi proteins relocate to the ER membrane. (C) By BFA removal, the punctate compartments with cis-Golgi proteins gather first to form the cis-cisternae. The Golgi stacksreassemble from the regenerated cis-cisternae in the cis-to-trans order (back to A).

6. RELATIONSHIP BETWEEN THE GOLGI APPARATUS AND THE TGN The TGN was first defined in mammalian cells as a specialized compartment on the trans-side of the Golgi apparatus composed of tubular networks (Griffiths and Simons, 1986; Roth et al., 1985; Taatjes and Roth, 1986). Similarly to the Golgi cisternae, the TGN also contains many resident enzymes for cargo processing (Rabouille et al., 1995). However, in addition to function as the sequel to the cargo processing line of the Golgi, the TGN sorts the cargo proteins destined for the plasma membrane and endosomes (Keller and Simons, 1997). Moreover, the TGN receives cargo from endosomal compartments (Pavelka et al., 1998). Thus, the TGN acts as the trafficking hub that both the secretory and endocytic pathways pass through. The Golgi cisternae and the TGN can be distinguished by the vesicles they form, because the Golgi cisternae generate COPI vesicles

266

Yoko Ito et al.

but not clathrin-coated vesicles, and the opposite is true for the TGN (Ladinsky et al., 2002). In plant cells, the TGN was first identified as “partially coated reticulum,” a vesiculotubular compartment with clathrin-coated buds (Pesacreta and Lucas, 1985), and later the word TGN was applied from mammalian studies (Staehelin et al., 1990). Similarly to the mammalian TGN, clathrin-coated vesicles are formed from the TGN, but not from the Golgi cisternae in plant cells (Donohoe et al., 2007). The plant TGN is marked by several SNARE proteins including SYP41, SYP61, and VPS45 (Uemura and Nakano, 2013). Using SYP41 and a subunit of V-ATPase (VHAa1) as the TGN markers, it has been demonstrated that a endocytosis tracer FM4–64 localizes to the TGN before late endosomes, indicating that the TGN functions as an early endosome (Dettmer et al., 2006). This is also supported by the observations by time-lapse imaging and immunoelectron microscopy, which have shown that the plasma membrane proteins internalized by endocytosis pass through the TGN (Viotti et al., 2010). In addition to the functional difference, dynamic behavior of the TGN is distinct from the Golgi cisternae. By electron tomography in Arabidopsis cells, the TGN seems to get separated from the Golgi (Staehelin and Kang, 2008). Such “free-floating TGN” is also reported in P. pastoris (Mogelsvang et al., 2003). Further observation by spinning disc confocal microscopy has revealed that the plant TGN moves independently of the Golgi apparatus and transiently associates with it (Viotti et al., 2010). Our detailed analysis by SCLIM has also demonstrated that there are two types of TGN; one is associated with the Golgi stacks and the other is independent from the Golgi (Uemura et al., in press). Such differences between the Golgi apparatus and the TGN may reflect their protein composition. Now it is possible to isolate and characterize the Golgi and TGN proteins separately. A recent study suggests that almost 30% of the proteins identified in the Arabidopsis TGN are non-Golgi proteins (Parsons et al., 2013). It is also known that the Golgi and the TGN can be distinguished by different behaviors upon BFA treatment. The formation of clathrin-coated vesicles at the TGN is thought to be regulated by ARF1, whose GEFs are affected by BFA. In mammalian cells, although the Golgi components are absorbed into the ER by BFA treatment, the TGN fuses with endosomes (Lippincott-Schwartz et al., 1991; Wood et al., 1991). In Arabidopsis root cells, in which the Golgi stacks do not disappear by BFA, the TGN forms large aggregates called BFA bodies (or BFA compartments) with endosomal

Golgi Apparatus of Plant Cells

267

components (Langhans et al., 2011; Uemura and Nakano, 2013). On the other hand, in tobacco cells, in which the Golgi stacks fuse and get absorbed into the ER upon BFA treatment, the TGN forms what look like BFA bodies, which are much smaller than those in Arabidopsis root cells (Langhans et al., 2011). This difference may be because the Golgi maturation continues and supplies materials to the TGN in the presence of BFA in Arabidopsis root cells but not in tobacco cells. Our detailed time-lapse observation tobacco BY-2 cells has also revealed that the regeneration process of the TGN after BFA removal is largely independent from the Golgi apparatus, but the association with the Golgi is indispensable (Ito et al., unpublished). These results indicate the complex nature of the plant TGN functioning both in the secretory and endocytic pathways.

7. CONCLUDING REMARKS As the application of electron microscopy promoted the Golgi apparatus from an artifact to the center of attention in cell biology, the improvement of light microscopy is bringing us the new perspective about the organelle being highly dynamic. Especially in these years, we are overcoming the diffraction limit of light microscopy by super-resolution microscopes. The plant Golgi apparatus provides an ideal system to apply such techniques, which will give us new insights about the structural organization and trafficking mechanisms of the Golgi stacks. Moreover, in spite of the high conservation of the stacked structure, the molecular components accomplishing the Golgi stacking appear to be different from organism to organism. The physiological significance of the rapid movement of the plant Golgi stacks throughout the cell is still mysterious. Further investigation of the plant Golgi apparatus including analysis of their cargo will shed light to the Golgi functions unique and important for plant life.

ACKNOWLEDGMENTS We thank members of the Nakano laboratory for valuable discussions. This work was supported by Grants-in-Aid for Scientific Research from the Ministry of Education, Culture, Sports, Science, and Technology of Japan and by funds from the Extreme Photonics and Cellular Systems Biology Projects of RIKEN. Y.I. is a recipient of the Research Fellowship for Young Scientists from the Japan Society for the Promotion of Science.

268

Yoko Ito et al.

REFERENCES Akkerman, M., Overdijk, E.J.R., Schel, J.H.N., Emons, A.M.C., Ketelaar, T., 2011. Golgi body motility in the plant cell cortex correlates with actin cytoskeleton organization. Plant Cell Physiol. 52, 1844–1855. Alcalde, J., Bonay, P., Roa, A., Vilaro, S., Sandoval, I.V., 1992. Assembly and disassembly of the Golgi complex: two processes arranged in a cis-trans direction. J. Cell Biol. 116, 69–83. Allan, B.B., Moyer, B.D., Balch, W.E., 2000. Rab1 recruitment of p115 into a cis-SNARE complex: programming budding COPII vesicles for fusion. Science 289, 444–448. Andag, U., Schmitt, H.D., 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. Andreeva, A.V., Zheng, H., Saint-Jore, C.M., Kutuzov, M.A., Evans, D.E., Hawes, C.R., 2000. Organization of transport from endoplasmic reticulum to Golgi in higher plants. Biochem. Soc. Trans. 28, 505–512. 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. Antonny, B., Madden, D., Hamamoto, S., Orci, L., Schekman, R., 2001. Dynamics of the COPII coat with GTP and stable analogues. Nat. Cell Biol. 3, 531–537. Appenzeller-Herzog, C., Hauri, H.-P., 2006. The ER-Golgi intermediate compartment (ERGIC): in search of its identity and function. J. Cell Sci. 119, 2173–2183. Avisar, D., Prokhnevsky, A.I., Makarova, K.S., Koonin, E.V., Dolja, V.V., 2008. Myosin XI-K is required for rapid trafficking of Golgi stacks, peroxisomes, and mitochondria in leaf cells of Nicotiana benthamiana. Plant Physiol. 146, 1098–1108. Avisar, D., Abu-Abied, M., Belausov, E., Sadot, E., 2012. Myosin XIK is a major player in cytoplasm dynamics and is regulated by two amino acids in its tail. J. Exp. Bot. 63, 241–249. Axelsson, M.A.B., Warren, G., 2004. Rapid, endoplasmic reticulum-independent diffusion of the mitotic Golgi haze. Mol. Cell. Biol. 15, 1843–1852. Bakker, H., Rouwendal, G.J.A., Karnoup, A.S., Florack, D.E.A., Stoopen, G.M., Helsper, J.P.F.G., van Ree, R., van Die, I., Bosch, D., 2006. An antibody produced in tobacco expressing a hybrid b-1,4-galactosyltransferase is essentially devoid of plant carbohydrate epitopes. Proc. Natl. Acad. Sci. U. S. A. 103, 7577–7582. Ballensiefen, W., Ossipov, D., Schmitt, H.D., 1998. Recycling of the yeast v-SNARE Sec22p involves COPI-proteins and the ER transmembrane proteins Ufe1p and Sec20p. J. Cell Sci. 111, 1507–1520. Bannykh, S.I., Rowe, T., Balch, W.E., 1996. The organization of endoplasmic reticulum export complexes. J. Cell Biol. 135, 19–35. Barlowe, C.K., Miller, E.A., 2013. Secretory protein biogenesis and traffic in the early secretory pathway. Genetics 193, 383–410. Barlowe, C., Schekman, R., 1993. SEC12 encodes a guanine-nucleotide-exchange factor essential for transport vesicle budding from the ER. Nature 365, 347–349. Barlowe, C., Orci, L., Yeung, T., Hosobuchi, M., Hamamoto, S., Salama, N., Rexach, M.F., 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. Barr, F.A., Puype, M., Vandekerckhove, J., Warren, G., 1997. GRASP65, a protein involved in the stacking of Golgi cisternae. Cell 91, 253–262. Barr, F.A., Nakamura, N., Warren, G., 1998. Mapping the interaction between GRASP65 and GM130, components of a protein complex involved in the stacking of Golgi cisternae. EMBO J. 17, 3258–3268.

Golgi Apparatus of Plant Cells

269

Barzilay, E., Ben-Califa, N., Hirschberg, K., Neumann, D., 2005. Uncoupling of brefeldin A-mediated coatomer protein complex-I dissociation from Golgi redistribution. Traffic 6, 794–802. Bassham, D.C., Brandizzi, F., Otegui, M.S., Sanderfoot, A.A., 2008. The secretory system of Arabidopsis. Arabidopsis Book 6, e0116. Batoko, H., Zheng, H.Q., Hawes, C., Moore, I., 2000. A Rab1 GTPase is required for transport between the endoplasmic reticulum and Golgi apparatus and for normal Golgi movement in plants. Plant Cell 12, 2201–2218. Beams, H.W., Kessel, R.G., 1968. The Golgi apparatus: structure and function. In: Bourne, G.H., Danielli, J.F., Jeon, K.W. (Eds.), Int. Rev. Cytol., vol. 23. Academic Press, New York, pp. 209–276. Beams, H.W., King, R.L., 1932. Cytoplasmic structures in the ganglion cells of certain Orthoptera, with special reference to the Golgi bodies, mitochondria, ‘vacuome’, intracellular trabeculae (trophospongium), and neurofibrillae. J. Morphol. 53, 59–95. Beams, H.W., King, R.L., 1935a. Effect of ultra-centrifuging on the cells of the root-tip of the bean. Nature 135, 232. Beams, H.W., King, R.L., 1935b. The effect of ultracentrifuging on the cells of the root tip of the bean (Phaseolus Vulgaris). Proc. R. Soc. Lond. Ser. B Biol. Sci. 118, 264–276. Beams, H.W., Tahmisian, T.N., Devine, R.L., Anderson, E., 1956. Electron microscope studies on the dictyosomes and acroblasts in the male germ cells of the cricket. J. Biophys. Biochem. Cytol. 2, 123–128. Beard, M., Satoh, A., Shorter, J., Warren, G., 2005. A cryptic Rab1-binding site in the p115 tethering protein. J. Biol. Chem. 280, 25840–25848. Becker, B., Bo¨linger, B., Melkonian, M., 1995. Anterograde transport of algal scales through the Golgi complex is not mediated by vesicles. Trends Cell Biol. 5, 305–307. Bednarek, S.Y., Reynolds, T.L., Schroeder, M., Grabowski, R., Hengst, L., Gallwitz, D., Raikhel, N.V., 1994. A small GTP-binding protein from Arabidopsis thaliana functionally complements the yeast YPT6 null mutant. Plant Physiol. 104, 591–596. Behnia, R., Barr, F.A., Flanagan, J.J., Barlowe, C., Munro, S., 2007. The yeast orthologue of GRASP65 forms a complex with a coiled-coil protein that contributes to ER to Golgi traffic. J. Cell Biol. 176, 255–261. Benchimol, M., Consort Ribeiro, K., Meyer Mariante, R., Alderete, J.F., 2001. Structure and division of the Golgi complex in Trichomonas vaginalis and Tritrichomonas foetus. Eur. J. Cell Biol. 80, 593–607. Benghezal, M., Wasteneys, G.O., Jones, D.A., 2000. The C-terminal dilysine motif confers endoplasmic reticulum localization to type I membrane proteins in plants. Plant Cell 12, 1179–1201. Ben-Tekaya, H., Miura, K., Pepperkok, R., Hauri, H.-P., 2005. Live imaging of bidirectional traffic from the ERGIC. J. Cell Sci. 118, 357–367. Bentivoglio, M., 1998. 1898: the Golgi apparatus emerges from nerve cells. Trends Neurosci. 21, 195–200. Berger, E., Rohrer, J., 2008. Golgi glycosylation enzymes. In: Mironov, A., Pavelka, M. (Eds.), The Golgi Apparatus. Springer, Vienna, pp. 161–189. Bevis, B.J., Hammond, A.T., Reinke, C.A., Glick, B.S., 2002. De novo formation of transitional ER sites and Golgi structures in Pichia pastoris. Nat. Cell Biol. 4, 750–756. Bhattacharyya, D., Glick, B.S., 2007. Two mammalian Sec16 homologues have nonredundant functions in endoplasmic reticulum (ER) export and transitional ER organization. Mol. Cell. Biol. 18, 839–849. Bi, X., Corpina, R.A., Goldberg, J., 2002. Structure of the Sec23/24-Sar1 pre-budding complex of the COPII vesicle coat. Nature 419, 271–277. Bi, X., Mancias, J.D., Goldberg, J., 2007. Insights into COPII coat nucleation from the structure of Sec23•Sar1 complexed with the active fragment of Sec31. Dev. Cell 13, 635–645.

270

Yoko Ito et al.

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. Blagitko, N., Schulz, U., Schinzel, A.A., Ropers, H.-H., Kalscheuer, V.M., 1999. g2-COP, a novel imprinted gene on chromosome 7q32, defines a new imprinting cluster in the human genome. Hum. Mol. Genet. 8, 2387–2396. Boevink, P., Oparka, K., Cruz, S.S., Martin, B., Betteridge, A., Hawes, C., 1998. Stacks on tracks: the plant Golgi apparatus traffics on an actin/ER network. Plant J. 15, 441–447. Bo¨hlenius, H., Mørch, S.M., Godfrey, D., Nielsen, M.E., Thordal-Christensen, H., 2010. The multivesicular body-localized GTPase ARFA1b/1c is important for callose deposition and ROR2 syntaxin-dependent preinvasive basal defense in barley. Plant Cell 22, 3831–3844. Boisson, M., Gomord, V., Audran, C., Berger, N., Dubreucq, B., Granier, F., Lerouge, P., Faye, L., Caboche, M., Lepiniec, L., 2001. Arabidopsis glucosidase I mutants reveal a critical role of N-glycan trimming in seed development. EMBO J. 20, 1010–1019. Bonfanti, L., Mironov Jr., A.A., Martı´nez-Mena´rguez, J.A., Martella, O., Fusella, A., Baldassarre, M., Buccione, R., Geuze, H.J., Mironov, A.A., Luini, A., 1998. Procollagen traverses the Golgi stack without leaving the lumen of cisternae: evidence for cisternal maturation. Cell 95, 993–1003. Boulaflous, A., Saint-Jore-Dupas, C., Herranz-Gordo, M.-C., Pagny-Salehabadi, S., Plasson, C., Garidou, F., Kiefer-Meyer, M.-C., Ritzenthaler, C., Faye, L., Gomord, V., 2009. Cytosolic N-terminal arginine-based signals together with a luminal signal target a type II membrane protein to the plant ER. BMC Plant Biol. 9, 144. Bowen, R.H., 1920. Studies on insect spermatogenesis: I. The history of the cytoplasmic components of the sperm in hemiptera. Biol. Bull. 39, 316–362, 311. Bowen, R.H., 1928. Studies on the structure of plant protoplasm. Z. Zellforsch. Mikrosk. Anat. 6, 689–725. Brandizzi, F., Barlowe, C., 2013. Organization of the ER-Golgi interface for membrane traffic control. Nat. Rev. Mol. Cell Biol. 14, 382–392. Brandizzi, F., Snapp, E.L., Roberts, A.G., Lippincott-Schwartz, J., Hawes, C., 2002. Membrane protein transport between the endoplasmic reticulum and the Golgi in tobacco leaves is energy dependent but cytoskeleton independent: evidence from selective photobleaching. Plant Cell 14, 1293–1309. Brown, R.M., Franke, W.W., Kleinig, H., Falk, H., Sitte, P., 1970. Scale formation in chrysophycean algae: I. Cellulosic and noncellulosic wall components made by the Golgi apparatus. J. Cell Biol. 45, 246–271. Brownhill, K., Wood, L., Allan, V., 2009. Molecular motors and the Golgi complex: staying put and moving through. Semin. Cell Dev. Biol. 20, 784–792. Bubeck, J., Scheuring, D., Hummel, E., Langhans, M., Viotti, C., Foresti, O., Denecke, J., Banfield, D.K., Robinson, D.G., 2008. The syntaxins SYP31 and SYP81 control ER– Golgi trafficking in the plant secretory pathway. Traffic 9, 1629–1652. Burn, J.E., Hurley, U.A., Birch, R.J., Arioli, T., Cork, A., Williamson, R.E., 2002. The cellulose-deficient Arabidopsis mutant rsw3 is defective in a gene encoding a putative glucosidase II, an enzyme processing N-glycans during ER quality control. Plant J. 32, 949–960. Burri, L., Varlamov, O., Doege, C.A., Hofmann, K., Beilharz, T., Rothman, J.E., So¨llner, T.H., Lithgow, T., 2003. A SNARE required for retrograde transport to the endoplasmic reticulum. Proc. Natl. Acad. Sci. U. S. A. 100, 9873–9877. Cai, H., Yu, S., Menon, S., Cai, Y., Lazarova, D., Fu, C., Reinisch, K., Hay, J.C., FerroNovick, S., 2007. TRAPPI tethers COPII vesicles by binding the coat subunit Sec23. Nature 445, 941–944. Campanoni, P., Blatt, M.R., 2007. Membrane trafficking and polar growth in root hairs and pollen tubes. J. Exp. Bot. 58, 65–74.

Golgi Apparatus of Plant Cells

271

Cao, X., Ballew, N., Barlowe, C., 1998. Initial docking of ER-derived vesicles requires Uso1p and Ypt1p but is independent of SNARE proteins. EMBO J. 17, 2156–2165. Caramelo, J.J., Parodi, A.J., 2008. Getting in and out from calnexin/calreticulin cycles. J. Biol. Chem. 283, 10221–10225. Cavalier-Smith, T., 1987. Eukaryotes with no mitochondria. Nature 326, 332–333. Ceriotti, A., Duranti, M., Bollini, R., 1998. Effects of N-glycosylation on the folding and structure of plant proteins. J. Exp. Bot. 49, 1091–1103. Chardin, P., McCormick, F., 1999. Brefeldin A: the advantage of being uncompetitive. Cell 97, 153–155. Chatre, L., Brandizzi, F., Hocquellet, A., Hawes, C., Moreau, P., 2005. Sec22 and Memb11 Are v-SNAREs of the anterograde endoplasmic reticulum-Golgi pathway in tobacco leaf epidermal cells. Plant Physiol. 139, 1244–1254. Chevalier, L., Bernard, S., Ramdani, Y., Lamour, R., Bardor, M., Lerouge, P., FolletGueye, M.-L., Driouich, A., 2010. Subcompartment localization of the side chain xyloglucan-synthesizing enzymes within Golgi stacks of tobacco suspension-cultured cells. Plant J. 64, 977–989. Choi, S.-w., Tamaki, T., Ebine, K., Uemura, T., Ueda, T., Nakano, A., 2013. RABA members act in distinct steps of subcellular trafficking of the FLAGELLIN SENSING2 receptor. Plant Cell 25, 1174–1187. Cluett, E.B., Brown, W.J., 1992. Adhesion of Golgi cisternae by proteinaceous interactions: intercisternal bridges as putative adhesive structures. J. Cell Sci. 103, 773–784. Conger, R., Chen, Y., Fornaciari, S., Faso, C., Held, M.A., Renna, L., Brandizzi, F., 2011. Evidence for the involvement of the Arabidopsis SEC24A in male transmission. J. Exp. Bot. 62, 4917–4926. Connerly, P.L., Esaki, M., Montegna, E.A., Strongin, D.E., Levi, S., Soderholm, J., Glick, B.S., 2005. Sec16 is a determinant of transitional ER organization. Curr. Biol. 15, 1439–1447. Contreras, I., Ortiz-Zapater, E., Aniento, F., 2004. Sorting signals in the cytosolic tail of membrane proteins involved in the interaction with plant ARF1 and coatomer. Plant J. 38, 685–698. Cosson, P., Amherdt, M., Rothman, J.E., Orci, L., 2002. A resident Golgi protein is excluded from peri-Golgi vesicles in NRK cells. Proc. Natl. Acad. Sci. U. S. A. 99, 12831–12834. Cox, K.M., Sterling, J.D., Regan, J.T., Gasdaska, J.R., Frantz, K.K., Peele, C.G., Black, A., Passmore, D., Moldovan-Loomis, C., Srinivasan, M., Cuison, S., Cardarelli, P.M., Dickey, L.F., 2006. Glycan optimization of a human monoclonal antibody in the aquatic plant Lemna minor. Nat. Biotechnol. 24, 1591–1597. Crowell, E.F., Bischoff, V., Desprez, T., Rolland, A., Stierhof, Y.-D., Schumacher, K., Gonneau, M., Ho¨fte, H., Vernhettes, S., 2009. Pausing of Golgi bodies on microtubules regulates secretion of cellulose synthase complexes in Arabidopsis. Plant Cell 21, 1141–1154. Dalton, A.J., 1951. Observations of the Golgi substance with the electron microscope. Nature 168, 244–245. Dalton, A.J., Felix, M.D., 1954. Cytologic and cytochemical characteristics of the Golgi substance of epithelial cells of the epididymis—in situ, in homogenates and after isolation. Am. J. Anat. 94, 171–207. daSilva, L.L., Snapp, E.L., Denecke, J., Lippincott-Schwartz, J., Hawes, C., Brandizzi, F., 2004. Endoplasmic reticulum export sites and Golgi bodies behave as single mobile secretory units in plant cells. Plant Cell 16, 1753–1771. de Graffenried, C.L., Bertozzi, C.R., 2004. The roles of enzyme localisation and complex formation in glycan assembly within the Golgi apparatus. Curr. Opin. Cell Biol. 16, 356–363.

272

Yoko Ito et al.

Denecke, J., De Rycke, R., Botterman, J., 1992. Plant and mammalian sorting signals for protein retention in the endoplasmic reticulum contain a conserved epitope. EMBO J. 11, 2345–2355. d’Enfert, C., Gensse, M., Gaillardin, C., 1992. Fission yeast and a plant have functional homologues of the Sar1 and Sec12 proteins involved in ER to Golgi traffic in budding yeast. EMBO J. 11, 4205–4211. Dettmer, J., Hong-Hermesdorf, A., Stierhof, Y.D., Schumacher, K., 2006. Vacuolar H+ATPase activity is required for endocytic and secretory trafficking in Arabidopsis. Plant Cell 18, 715–730. Diao, A., Rahman, D., Pappin, D.J.C., Lucocq, J., Lowe, M., 2003. The coiled-coil membrane protein golgin-84 is a novel rab effector required for Golgi ribbon formation. J. Cell Biol. 160, 201–212. Dilcher, M., Veith, B., Chidambaram, S., Hartmann, E., Schmitt, H.D., Fischer von Mollard, G., 2003. Use1p is a yeast SNARE protein required for retrograde traffic to the ER. EMBO J. 22, 3664–3674. Donaldson, J.G., Cassel, D., Kahn, R.A., Klausner, R.D., 1992a. ADP-ribosylation factor, a small GTP-binding protein, is required for binding of the coatomer protein beta-COP to Golgi membranes. Proc. Natl. Acad. Sci. U. S. A. 89, 6408–6412. Donaldson, J.G., Finazzi, D., Klausner, R.D., 1992b. Brefeldin A inhibits Golgi membranecatalysed exchange of guanine nucleotide onto ARF protein. Nature 360, 350–352. Donohoe, B.S., Kang, B.-H., Staehelin, L.A., 2007. Identification and characterization of COPIa- and COPIb-type vesicle classes associated with plant and algal Golgi. Proc. Natl. Acad. Sci. U. S. A. 104, 163–168. Donohoe, B.S., Kang, B.-H., Gerl, M.J., Gergely, Z.R., McMichael, C.M., Bednarek, S.Y., Staehelin, L.A., 2013. Cis-Golgi cisternal assembly and biosynthetic activation occur sequentially in plants and algae. Traffic 14, 551–567. Driouich, A., Follet-Gueye, M.-L., Bernard, S., Kousar, S., Chevalier, L., Vicre´, M., Lerouxel, O., 2012. Golgi-mediated synthesis and secretion of matrix polysaccharides of the primary cell wall of higher plants. Front. Plant. Sci. 3, 79. Dunphy, W.G., Rothman, J.E., 1985. Compartmental organization of the Golgi stack. Cell 42, 13–21. Dupree, P., Sherrier, D.J., 1998. The plant Golgi apparatus. Mol. Cell. Res. 1404, 259–270, BBA. Egea, G., Rı´os, R., 2008. The role of the cytoskeleton in the structure and function of the Golgi apparatus. In: Mironov, A., Pavelka, M. (Eds.), The Golgi Apparatus. Springer, Vienna, pp. 270–300. Egea, G., Serra-Peinado, C., Salcedo-Sicilia, L., Gutie´rrez-Martı´nez, E., 2013. Actin acting at the Golgi. Histochem. Cell Biol. 140, 347–360. Espenshade, P., Gimeno, R.E., Holzmacher, E., Teung, P., Kaiser, C.A., 1995. Yeast SEC16 gene encodes a multidomain vesicle coat protein that interacts with Sec23p. J. Cell Biol. 131, 311–324. Farid, A., Malinovsky, F.G., Veit, C., Schoberer, J., Zipfel, C., Strasser, R., 2013. Specialized roles of the conserved subunit OST3/6 of the oligosaccharyltransferase complex in innate immunity and tolerance to abiotic stresses. Plant Physiol. 162, 24–38. Faso, C., Chen, Y.N., Tamura, K., Held, M., Zemelis, S., Marti, L., Saravanan, R.S., Hummel, E., Kung, L., Miller, E., Hawes, C., Brandizzi, F., 2009. A missense mutation in the Arabidopsis COPII coat protein Sec24A induces the formation of clusters of the endoplasmic reticulum and Golgi apparatus. Plant Cell 21, 3655–3671. Feinstein, T.N., Linstedt, A.D., 2008. GRASP55 regulates Golgi ribbon formation. Mol. Cell. Biol. 19, 2696–2707. Franke, W.W., Kartenbeck, J., Krien, S., Woude, W.J., Scheer, U., Morre´, D.J., 1972. Interand intracisternal elements of the Golgi apparatus. Z. Zellforsch. Mikrosk. Anat. 132, 365–380.

Golgi Apparatus of Plant Cells

273

Fukuda, M., Kanno, E., Ishibashi, K., Itoh, T., 2008. Large scale screening for novel Rab effectors reveals unexpected broad Rab binding specificity. Mol. Cell. Proteomics 7, 1031–1042. Fu¨llekrug, J., Boehm, J., Ro¨ttger, S., Nilsson, T., Mieskes, G., Schmitt, H.D., 1997a. Human Rer1 is localized to the Golgi apparatus and complements the deletion of the homologous Rer1 protein of Saccharomyces cerevisiae. Eur. J. Cell Biol. 74, 31–40. Fu¨llekrug, J., So¨nnichsen, B., Scha¨fer, U., Nguyen Van, P., So¨ling, H.D., Mieskes, G., 1997b. Characterization of brefeldin A induced vesicular structures containing cycling proteins of the intermediate compartment/cis-Golgi network. FEBS Lett. 404, 75–81. Futatsumori, M., Kasai, K., Takatsu, H., Shin, H.-W., Nakayama, K., 2000. Identification and characterization of novel isoforms of COP I subunits. J. Biochem. 128, 793–801. Garcia-Herdugo, G., Gonza´lez-Reyes, J.A., Gracia-Navarro, F., Navas, P., 1988. Growth kinetics of the Golgi apparatus during the cell cycle in onion root meristems. Planta 175, 305–312. Geldner, N., Anders, N., Wolters, H., Keicher, J., Kornberger, W., Muller, P., Delbarre, A., Ueda, T., Nakano, A., Ju¨rgens, G., 2003. The Arabidopsis GNOM ARF-GEF mediates endosomal recycling auxin transport, and auxin-dependent plant g. Cell 112, 219–230. Gilchrist, A., Au, C.E., Hiding, J., Bell, A.W., Fernandez-Rodriguez, J., Lesimple, S., Nagaya, H., Roy, L., Gosline, S.J.C., Hallett, M., Paiement, J., Kearney, Robert E., Nilsson, T., Bergeron, J.J.M., 2006. Quantitative proteomics analysis of the secretory pathway. Cell 127, 1265–1281. Gimeno, R.E., Espenshade, P., Kaiser, C.A., 1996. COPII coat subunit interactions: Sec24p and Sec23p bind to adjacent regions of Sec16p. Mol. Cell. Biol. 7, 1815–1823. Girod, A., Storrie, B., Simpson, J.C., Johannes, L., Goud, B., Roberts, L.M., Lord, J.M., Nilsson, T., Pepperkok, R., 1999. Evidence for a COP-I-independent transport route from the Golgi complex to the endoplasmic reticulum. Nat. Cell Biol. 1, 423–430. Glick, B.S., Luini, A., 2011. Models for Golgi traffic: a critical assessment. Cold Spring Harb. Perspect. Biol. 3, a005215. Glick, B.S., Nakano, A., 2009. Membrane traffic within the Golgi apparatus. Annu. Rev. Cell Dev. Biol. 25, 113–132. Goldberg, J., 1998. Structural basis for activation of ARF GTPase: mechanisms of guanine nucleotide exchange and GTP myristoyl switching. Cell 95, 237–248. Goldberg, J., 2000. Decoding of sorting signals by coatomer through a GTPase switch in the COPI coat complex. Cell 100, 671–679. Golgi, C., 1898. Sur la structure des cellules nerveuses. Arch. Ital. Biol. 30, 60–71. Gomord, V., Fitchette, A.-C., Menu-Bouaouiche, L., Saint-Jore-Dupas, C., Plasson, C., Michaud, D., Faye, L., 2010. Plant-specific glycosylation patterns in the context of therapeutic protein production. Plant Biotechnol. J. 8, 564–587. Grabenhorst, E., Conradt, H.S., 1999. The cytoplasmic, transmembrane, and stem regions of glycosyltransferases specify their in vivo functional sublocalization and stability in the Golgi. J. Biol. Chem. 274, 36107–36116. Grasse´, P.P., 1957. Ultrastructure, polarite´ et reproduction de l’appareil de Golgi. Compt. Rend. Acad. Sci. 245, 1278–1281. Griffiths, G., Simons, K., 1986. The trans Golgi network: sorting at the exit site of the Golgi complex. Science 234, 438–443. Guo, Y., Punj, V., Sengupta, D., Linstedt, A.D., 2008. Coat-tether interaction in Golgi organization. Mol. Cell. Biol. 19, 2830–2843. Gutierrez, R., Lindeboom, J.J., Paredez, A.R., Emons, A.M.C., Ehrhardt, D.W., 2009. Arabidopsis cortical microtubules position cellulose synthase delivery to the plasma membrane and interact with cellulose synthase trafficking compartments. Nat. Cell Biol. 11, 797–806. Hamada, T., Tominaga, M., Fukaya, T., Nakamura, M., Nakano, A., Watanabe, Y., Hashimoto, T., Baskin, T.I., 2012. RNA processing bodies, peroxisomes, Golgi bodies,

274

Yoko Ito et al.

mitochondria, and endoplasmic reticulum tubule junctions frequently pause at cortical microtubules. Plant Cell Physiol. 53, 699–708. Hammond, A.T., Glick, B.S., 2000. Dynamics of transitional endoplasmic reticulum sites in vertebrate cells. Mol. Cell. Biol. 11, 3013–3030. Hanton, S.L., Renna, L., Bortolotti, L.E., Chatre, L., Stefano, G., Brandizzi, F., 2005. Diacidic motifs influence the export of transmembrane proteins from the endoplasmic reticulum in plant cells. Plant Cell 17, 3081–3093. Hanton, S.L., Chatre, L., Renna, L., Matheson, L.A., Brandizzi, F., 2007. De novo formation of plant endoplasmic reticulum export sites is membrane cargo induced and signal mediated. Plant Physiol. 143, 1640–1650. Hanton, S., Chatre, L., Matheson, L., Rossi, M., Held, M., Brandizzi, F., 2008. Plant Sar1 isoforms with near-identical protein sequences exhibit different localisations and effects on secretion. Plant Mol. Biol. 67, 283–294. Hanton, S.L., Matheson, L.A., Chatre, L., Brandizzi, F., 2009. Dynamic organization of COPII coat proteins at endoplasmic reticulum export sites in plant cells. Plant J. 57, 963–974. Ha¨weker, H., Rips, S., Koiwa, H., Salomon, S., Saijo, Y., Chinchilla, D., Robatzek, S., von Schaewen, A., 2010. Pattern recognition receptors require N-glycosylation to mediate plant immunity. J. Biol. Chem. 285, 4629–4636. Hawes, C., 2012. The ER/Golgi interface—is there anything in-between? Front. Plant. Sci. 3, 73. Hawes, C., Osterrieder, A., Hummel, E., Sparkes, I., 2008a. The plant ER-Golgi interface. Traffic 9, 1571–1580. Hawes, C., Osterrieder, A., Sparkes, I., 2008b. Features of the plant Golgi apparatus. In: Mironov, A., Pavelka, M. (Eds.), The Golgi Apparatus. Springer, Vienna, pp. 611–622. He, C.Y., Ho, H.H., Malsam, J., Chalouni, C., West, C.M., Ullu, E., Toomre, D., Warren, G., 2004. Golgi duplication in Trypanosoma brucei. J. Cell Biol. 165, 313–321. Helms, J.B., Rothman, J.E., 1992. Inhibition by brefeldin A of a Golgi membrane enzyme that catalyses exchange of guanine nucleotide bound to ARF. Nature 360, 352–354. Hepler, P.K., Wolniak, S.M., 1984. Membranes in the Mitotic apparatus: their structure and function. In: Bourne, G.H., Danielli, J.F., Jeon, K.W. (Eds.), Int. Rev. Cytol., vol. 90. Academic Press, New York, pp. 169–238. Herpers, B., Rabouille, C., 2004. mRNA localization and ER-based protein sorting mechanisms dictate the use of transitional endoplasmic reticulum-Golgi units involved in Gurken transport in Drosophila Oocytes. Mol. Cell. Biol. 15, 5306–5317. Hirose, S., Komamine, A., 1989. Changes in ultrastructure of Golgi apparatus during the cell cycle in a synchronous culture of Catharanthus roseus. New Phytol. 111, 599–605. Ho, H.H., He, C.Y., de Graffenried, C.L., Murrells, L.J., Warren, G., 2006. Ordered assembly of the duplicating Golgi in Trypanosoma brucei. Proc. Natl. Acad. Sci. U. S. A. 103, 7676–7681. Horton, A.C., Ehlers, M.D., 2003. Dual modes of endoplasmic reticulum-to-Golgi transport in dendrites revealed by live-cell imaging. J. Neurosci. 23, 6188–6199. Hsu, V., Yang, J.-S., Lee, S., 2008. COPI: mechanisms and transport roles. In: Mironov, A., Pavelka, M. (Eds.), The Golgi Apparatus. Springer, Vienna, pp. 87–105. Huang, M., Weissman, J.T., Be´raud-Dufour, S., Luan, P., Wang, C., Chen, W., Aridor, M., Wilson, I.A., Balch, W.E., 2001. Crystal structure of Sar1-GDP at 1.7 A˚ resolution and the role of the NH2 terminus in ER export. J. Cell Biol. 155, 937–948. Hurtado, L., Caballero, C., Gavilan, M.P., Cardenas, J., Bornens, M., Rios, R.M., 2011. Disconnecting the Golgi ribbon from the centrosome prevents directional cell migration and ciliogenesis. J. Cell Biol. 193, 917–933. Igdoura, S.A., Herscovics, A., Lal, A., Moremen, K.W., Morales, C.R., Hermo, L., 1999. a-Mannosidases involved in N-glycan processing show cell specificity and distinct

Golgi Apparatus of Plant Cells

275

subcompartmentalization within the Golgi apparatus of cells in the testis and epididymis. Eur. J. Cell Biol. 78, 441–452. Ioffe, E., Stanley, P., 1994. Mice lacking N-acetylglucosaminyltransferase I activity die at mid-gestation, revealing an essential role for complex or hybrid N-linked carbohydrates. Proc. Natl. Acad. Sci. U. S. A. 91, 728–732. Ishikawa, T., Machida, C., Yoshioka, Y., Ueda, T., Nakano, A., Machida, Y., 2008. EMBRYO YELLOW gene, encoding a subunit of the conserved oligomeric Golgi complex, is required for appropriate cell expansion and meristem organization in Arabidopsis thaliana. Genes Cells 13, 521–535. Ito, Y., Uemura, T., Shoda, K., Fujimoto, M., Ueda, T., Nakano, A., 2012. cis-Golgi proteins accumulate near the ER exit sites and act as the scaffold for Golgi regeneration after brefeldin A treatment in tobacco BY-2 cells. Mol. Cell. Biol. 23, 3203–3214. Ivan, V., de Voer, G., Xanthakis, D., Spoorendonk, K.M., Kondylis, V., Rabouille, C., 2008. Drosophila Sec16 mediates the biogenesis of tER sites upstream of Sar1 through an arginine-rich motif. Mol. Cell. Biol. 19, 4352–4365. Jackson, C.L., Casanova, J.E., 2000. Turning on ARF: the Sec7 family of guaninenucleotide-exchange factors. Trends Cell Biol. 10, 60–67. Jamieson, J.D., Palade, G.E., 1966. Role of the Golgi complex in the intracellular transport of secretory proteins. Proc. Natl. Acad. Sci. U. S. A. 55, 424–431. Je´kely, G., 2008. Evolution of the Golgi complex. In: Mironov, A., Pavelka, M. (Eds.), The Golgi Apparatus. Springer, Vienna, pp. 675–691. Ju¨rgens, G., 2004. Membrane trafficking in plants. Annu. Rev. Cell Dev. Biol. 20, 481–504. Kang, B.-H., Staehelin, L.A., 2008. ER-to-Golgi transport by COPII vesicles in Arabidopsis involves a ribosome-excluding scaffold that is transferred with the vesicles to the Golgi matrix. Protoplasma 234, 51–64. Kang, J.S., Frank, J., Kang, C.H., Kajiura, H., Vikram, M., Ueda, A., Kim, S., Bahk, J.D., Triplett, B., Fujiyama, K., Lee, S.Y., von Schaewen, A., Koiwa, H., 2008. Salt tolerance of Arabidopsis thaliana requires maturation of N-glycosylated proteins in the Golgi apparatus. Proc. Natl. Acad. Sci. U. S. A. 105, 5933–5938. Keller, P., Simons, K., 1997. Post-Golgi biosynthetic trafficking. J. Cell Sci. 110, 3001–3009. Kim, W., Cheong, N., Lee, D., Lee, K., Je, D., Bahk, J., Cho, M., Lee, S., 1996. Isolation of an additional soybean cDNA encoding Ypt/Rab-related small GTP-binding protein and its functional comparison to Sypt using a yeast ypt1-1 mutant. Plant Mol. Biol. 31, 783–792. Kloepper, T.H., Kienle, C.N., Fasshauer, D., 2007. An elaborate classification of SNARE proteins sheds light on the conservation of the eukaryotic endomembrane system. Mol. Cell. Biol. 18, 3463–3471. Klute, M.J., Melanc¸on, P., Dacks, J.B., 2011. Evolution and diversity of the Golgi. Cold Spring Harb. Perspect. Biol. 3, a007849. Kondylis, V., Rabouille, C., 2009. The Golgi apparatus: lessons from Drosophila. FEBS Lett. 583, 3827–3838. Kondylis, V., Spoorendonk, K.M., Rabouille, C., 2005. dGRASP localization and function in the early exocytic pathway in Drosophila S2 Cells. Mol. Cell. Biol. 16, 4061–4072. Kraynack, B.A., Chan, A., Rosenthal, E., Essid, M., Umansky, B., Waters, M.G., Schmitt, H.D., 2005. Dsl1p, Tip20p, and the novel Dsl3(Sec39) protein are required for the stability of the Q/t-SNARE complex at the endoplasmic reticulum in yeast. Mol. Cell. Biol. 16, 3963–3977. Kuehn, M.J., Herrmann, J.M., Schekman, R., 1998. COPII-cargo interactions direct protein sorting into ER-derived transport vesicles. Nature 391, 187–190. Kuge, O., Dascher, C., Orci, L., Rowe, T., Amherdt, M., Plutner, H., Ravazzola, M., Tanigawa, G., Rothman, J.E., Balch, W.E., 1994. Sar1 promotes vesicle budding from the endoplasmic reticulum but not Golgi compartments. J. Cell Biol. 125, 51–65.

276

Yoko Ito et al.

Kurihara, T., Hamamoto, S., Gimeno, R.E., Kaiser, C.A., Schekman, R., Yoshihisa, T., 2000. Sec24p and Iss1p function interchangeably in transport vesicle formation from the endoplasmic reticulum in Saccharomyces cerevisiae. Mol. Cell. Biol. 11, 983–998. Kweon, H.-S., Beznoussenko, G.V., Micaroni, M., Polishchuk, R.S., Trucco, A., Martella, O., Di Giandomenico, D., Marra, P., Fusella, A., Di Pentima, A., Berger, E.G., Geerts, W.J.C., Koster, A.J., Burger, K.N.J., Luini, A., Mironov, A.A., 2004. Golgi enzymes are enriched in perforated zones of Golgi cisternae but are depleted in COPI vesicles. Mol. Cell. Biol. 15, 4710–4724. Ladinsky, M.S., Wu, C.C., McIntosh, S., McIntosh, J.R., Howell, K.E., 2002. Structure of the Golgi and distribution of reporter molecules at 20 C reveals the complexity of the exit compartments. Mol. Cell. Biol. 13, 2810–2825. Langhans, M., Hawes, C., Hillmer, S., Hummel, E., Robinson, D.G., 2007. Golgi regeneration after brefeldin A treatment in BY-2 cells entails stack enlargement and cisternal growth followed by division. Plant Physiol. 145, 527–538. Langhans, M., Marcote, M.J., Pimpl, P., Virgili-Lo´pez, G., Robinson, D.G., Aniento, F., 2008. In vivo trafficking and localization of p24 proteins in plant cells. Traffic 9, 770–785. Langhans, M., Fo¨rster, S., Helmchen, G., Robinson, D.G., 2011. Differential effects of the brefeldin A analogue (6R)-hydroxy-BFA in tobacco and Arabidopsis. J. Exp. Bot. 62, 2949–2957. Langhans, M., Meckel, T., Kress, A., Lerich, A., Robinson, D., 2012. ERES (ER exit sites) and the “secretory unit concept” J. Microsc. 247, 48–59. Latijnhouwers, M., Hawes, C., Carvalho, C., 2005. Holding it all together? Candidate proteins for the plant Golgi matrix. Curr. Opin. Plant Biol. 8, 632–639. Latijnhouwers, M., Gillespie, T., Boevink, P., Kriechbaumer, V., Hawes, C., Carvalho, C.M., 2007. Localization and domain characterization of Arabidopsis golgin candidates. J. Exp. Bot. 58, 4373–4386. Lavieu, G., Zheng, H., Rothman, J.E., 2013. Stapled Golgi cisternae remain in place as cargo passes through the stack. ELife 2, e00558. Lee, H.I., Gal, S., Newman, T.C., Raikhel, N.V., 1993. The Arabidopsis endoplasmic reticulum retention receptor functions in yeast. Proc. Natl. Acad. Sci. U. S. A. 90, 11433–11437. Lee, S.Y., Yang, J.-S., Hong, W., Premont, R.T., Hsu, V.W., 2005. ARFGAP1 plays a central role in coupling COPI cargo sorting with vesicle formation. J. Cell Biol. 168, 281–290. Lerich, A., Hillmer, S., Langhans, M., Scheuring, D., van Bentum, P., Robinson, D.G., 2012. ER import sites and their relationship to ER exit sites: a new model for bidirectional ER-Golgi transport in higher plants. Front. Plant. Sci. 3, 143. Letourneur, F., Gaynor, E.C., Hennecke, S., De´mollie`re, C., Duden, R., Emr, S.D., Riezman, H., Cosson, P., 1994. Coatomer is essential for retrieval of dilysine-tagged proteins to the endoplasmic reticulum. Cell 79, 1199–1207. Levi, S.K., Bhattacharyya, D., Strack, R.L., Austin, J.R., Glick, B.S., 2010. The yeast GRASP Grh1 colocalizes with COPII and is dispensable for organizing the secretory pathway. Traffic 11, 1168–1179. Lewis, M.J., Pelham, H.R., 1990. A human homologue of the yeast HDEL receptor. Nature 348, 162–163. Lewis, M.J., Pelham, H.R.B., 1996. SNARE-mediated retrograde traffic from the Golgi complex to the endoplasmic reticulum. Cell 85, 205–215. Lewis, M.J., Sweet, D.J., Pelham, H.R.B., 1990. The ERD2 gene determines the specificity of the luminal ER protein retention system. Cell 61, 1359–1363. Li, L., Shimada, T., Takahashi, H., Ueda, H., Fukao, Y., Kondo, M., Nishimura, M., HaraNishimura, I., 2006. MAIGO2 is involved in exit of seed storage proteins from the endoplasmic reticulum in Arabidopsis thaliana. Plant Cell 18, 3535–3547.

Golgi Apparatus of Plant Cells

277

Li, L., Shimada, T., Takahashi, H., Koumoto, Y., Shirakawa, M., Takagi, J., Zhao, X., Tu, B., Jin, H., Shen, Z., Han, B., Jia, M., Kondo, M., Nishimura, M., Hara-Nishimura, I., 2013. MAG2 and three MAG2 INTERACTING PROTEINs form an ER-localized complex to facilitate the storage protein transport in Arabidopsis thaliana. Plant J. 76, 781–791. Liebminger, E., Hu¨ttner, S., Vavra, U., Fischl, R., Schoberer, J., Grass, J., Blaukopf, C., Seifert, G.J., Altmann, F., Mach, L., Strasser, R., 2009. Class I a-Mannosidases are required for N-glycan processing and root development in Arabidopsis thaliana. Plant Cell 21, 3850–3867. Lige, B., Ma, S., van Huystee, R.B., 2001. The effects of the site-directed removal of N-glycosylation from cationic peanut peroxidase on its function. Arch. Biochem. Biophys. 386, 17–24. Lippincott-Schwartz, J., Yuan, L.C., Bonifacino, J.S., Klausner, R.D., 1989. Rapid redistribution of Golgi proteins into the ER in cells treated with brefeldin A: evidence for membrane cycling from Golgi to ER. Cell 56, 801–813. Lippincott-Schwartz, J., Donaldson, J.G., Schweizer, A., Berger, E.G., Hauri, H.P., Yuan, L.C., Klausner, R.D., 1990. Microtubule-dependent retrograde transport of proteins into the ER in the presence of brefeldin A suggests an ER recycling pathway. Cell 60, 821–836. Lippincott-Schwartz, J., Yuan, L., Tipper, C., Amherdt, M., Orci, L., Klausner, R.D., 1991. Brefeldin A’s effects on endosomes, lysosomes, and the TGN suggest a general mechanism for regulating organelle structure and membrane traffic. Cell 67, 601–616. Liu, J.-X., Howell, S.H., 2010. Endoplasmic reticulum protein quality control and its relationship to environmental stress responses in plants. Plant Cell 22, 2930–2942. Losev, E., Reinke, C.A., Jellen, J., Strongin, D.E., Bevis, B.J., Glick, B.S., 2006. Golgi maturation visualized in living yeast. Nature 441, 1002–1006. Lowe, M., 2011. Structural organization of the Golgi apparatus. Curr. Opin. Cell Biol. 23, 85–93. Luo, L., Hannemann, M., Koenig, S., Hegermann, J., Ailion, M., Cho, M.-K., Sasidharan, N., Zweckstetter, M., Rensing, S.A., Eimer, S., 2011. The Caenorhabditis elegans GARP complex contains the conserved Vps51 subunit and is required to maintain lysosomal morphology. Mol. Cell. Biol. 22, 2564–2578. Madison, S.L., Nebenfu¨hr, A., 2011. Live-cell imaging of dual-labeled Golgi stacks in tobacco BY-2 cells reveals similar behaviors for different cisternae during movement and Brefeldin A treatment. Mol. Plant 4, 896–908. Malsam, J., Satoh, A., Pelletier, L., Warren, G., 2005. Golgin tethers define subpopulations of COPI vesicles. Science 307, 1095–1098. Manolea, F., Claude, A., Chun, J., Rosas, J., Melanc¸on, P., 2008. Distinct functions for Arf guanine nucleotide exchange factors at the Golgi complex: GBF1 and BIGs are required for assembly and maintenance of the Golgi stack and trans-Golgi network, respectively. Mol. Cell. Biol. 19, 523–535. Marra, P., Salvatore, L., Mironov, A., Di Campli, A., Di Tullio, G., Trucco, A., Beznoussenko, G., De Matteis, M.A., 2007. The biogenesis of the Golgi ribbon: the roles of membrane input from the ER and of GM130. Mol. Cell. Biol. 18, 1595–1608. Martı´nez-Mena´rguez, J.A., Geuze, H.J., Slot, J.W., Klumperman, J., 1999. Vesicular tubular clusters between the ER and Golgi mediate concentration of soluble secretory proteins by exclusion from COPI-coated vesicles. Cell 98, 81–90. Martı´nez-Mena´rguez, J.A., Prekeris, R., Oorschot, V.M.J., Scheller, R., Slot, J.W., Geuze, H.J., Klumperman, J., 2001. Peri-Golgi vesicles contain retrograde but not anterograde proteins consistent with the cisternal progression model of intra-Golgi transport. J. Cell Biol. 155, 1213–1224.

278

Yoko Ito et al.

Matsuoka, K., Schekman, R., Orci, L., Heuser, J.E., 2001. Surface structure of the COPIIcoated vesicle. Proc. Natl. Acad. Sci. U. S. A. 98, 13705–13709. Matsuura-Tokita, K., Takeuchi, M., Ichihara, A., Mikuriya, K., Nakano, A., 2006. Live imaging of yeast Golgi cisternal maturation. Nature 441, 1007–1010. Matteis, M., Mironov, A., Beznoussenko, G., 2008. The Golgi ribbon and the function of the Golgins. In: Mironov, A., Pavelka, M. (Eds.), The Golgi Apparatus. Springer, Vienna, pp. 223–246. McCartney, A.W., Dyer, J.M., Dhanoa, P.K., Kim, P.K., Andrews, D.W., McNew, J.A., Mullen, R.T., 2004. Membrane-bound fatty acid desaturases are inserted co-translationally into the ER and contain different ER retrieval motifs at their carboxy termini. Plant J. 37, 156–173. Melanc¸on, P., Glick, B.S., Malhotra, V., Weidman, P.J., Serafini, T., Gleason, M.L., Orci, L., Rothman, J.E., 1987. Involvement of GTP-binding G proteins in transport through the Golgi stack. Cell 51, 1053–1062. Melkonian, M., Becker, B., Becker, D., 1991. Scale formation in algae. J. Electron Microsc. Tech. 17, 165–178. Mellman, I., Warren, G., 2000. The road taken: past and future foundations of membrane traffic. Cell 100, 99–112. Memon, A.R., 2004. The role of ADP-ribosylation factor and SAR1 in vesicular trafficking in plants. Biochim. Biophys. Acta 1664, 9–30. Merte, J., Jensen, D., Wright, K., Sarsfield, S., Wang, Y., Schekman, R., Ginty, D.D., 2010. Sec24b selectively sorts Vangl2 to regulate planar cell polarity during neural tube closure. Nat. Cell Biol. 12, 41–46. Metzler, M., Gertz, A., Sarkar, M., Schachter, H., Schrader, J.W., Marth, J.D., 1994. Complex asparagine-linked oligosaccharides are required for morphogenic events during post-implantation development. EMBO J. 13, 2056–2065. Michelsen, K., Schmid, V., Metz, J., Heusser, K., Liebel, U., Schwede, T., Spang, A., Schwappach, B., 2007. Novel cargo-binding site in the b and d subunits of coatomer. J. Cell Biol. 179, 209–217. Miller, E.A., Beilharz, T.H., Malkus, P.N., Lee, M.C.S., Hamamoto, S., Orci, L., Schekman, R., 2003. Multiple cargo binding sites on the COPII subunit Sec24p ensure capture of diverse membrane proteins into transport vesicles. Cell 114, 497–509. Min, M.K., Kim, S.J., Miao, Y., Shin, J., Jiang, L., Hwang, I., 2007. Overexpression of Arabidopsis AGD7 causes relocation of Golgi-localized proteins to the endoplasmic reticulum and inhibits protein trafficking in plant cells. Plant Physiol. 143, 1601–1614. Min, M.K., Jang, M., Lee, M., Lee, J., Song, K., Lee, Y., Choi, K.Y., Robinson, D.G., Hwang, I., 2013. Recruitment of Arf1-GDP to Golgi by Glo3p-Type ArfGAPs is crucial for Golgi maintenance and plant growth. Plant Physiol. 161, 676–691. Mironov, A.A., Beznoussenko, G.V., Nicoziani, P., Martella, O., Trucco, A., Kweon, H.-S., Di Giandomenico, D., Polishchuk, R.S., Fusella, A., Lupetti, P., Berger, E.G., Geerts, W.J.C., Koster, A.J., Burger, K.N.J., Luini, A., 2001. Small cargo proteins and large aggregates can traverse the Golgi by a common mechanism without leaving the lumen of cisternae. J. Cell Biol. 155, 1225–1238. Moelleken, J., Malsam, J., Betts, M.J., Movafeghi, A., Reckmann, I., Meissner, I., Hellwig, A., Russell, R.B., So¨llner, T., Bru¨gger, B., Wieland, F.T., 2007. Differential localization of coatomer complex isoforms within the Golgi apparatus. Proc. Natl. Acad. Sci. U. S. A. 104, 4425–4430. Mogelsvang, S., Gomez-Ospina, N., Soderholm, J., Glick, B.S., Staehelin, L.A., 2003. Tomographic evidence for continuous turnover of Golgi cisternae in Pichia pastoris. Mol. Cell. Biol. 14, 2277–2291.

Golgi Apparatus of Plant Cells

279

Mollenhauer, H.H., 1965. An intercisternal structure in the Golgi apparatus. J. Cell Biol. 24, 504–511. Montesinos, J.C., Sturm, S., Langhans, M., Hillmer, S., Marcote, M.J., Robinson, D.G., Aniento, F., 2012. Coupled transport of Arabidopsis p24 proteins at the ER–Golgi interface. J. Exp. Bot. 63, 4243–4261. Moore, P.J., Swords, K.M., Lynch, M.A., Staehelin, L.A., 1991. Spatial organization of the assembly pathways of glycoproteins and complex polysaccharides in the Golgi apparatus of plants. J. Cell Biol. 112, 589–602. Moremen, K.W., Tiemeyer, M., Nairn, A.V., 2012. Vertebrate protein glycosylation: diversity, synthesis and function. Nat. Rev. Mol. Cell Biol. 13, 448–462. Mossessova, E., Bickford, L.C., Goldberg, J., 2003. SNARE selectivity of the COPII coat. Cell 114, 483–495. Mowbrey, K., Dacks, J.B., 2009. Evolution and diversity of the Golgi body. FEBS Lett. 583, 3738–3745. Munro, S., 1995. An investigation of the role of transmembrane domains in Golgi protein retention. EMBO J. 14, 4695–4704. Nakamura, N., Rabouille, C., Watson, R., Nilsson, T., Hui, N., Slusarewicz, P., Kreis, T.E., Warren, G., 1995. Characterization of a cis-Golgi matrix protein, GM130. J. Cell Biol. 131, 1715–1726. Nakano, A., Luini, A., 2010. Passage through the Golgi. Curr. Opin. Cell Biol. 22, 471–478. Nakano, A., Muramatsu, M., 1989. A novel GTP-binding protein, Sar1p, is involved in transport from the endoplasmic reticulum to the Golgi apparatus. J. Cell Biol. 109, 2677–2691. Nakano, R.T., Matsushima, R., Ueda, H., Tamura, K., Shimada, T., Li, L., Hayashi, Y., Kondo, M., Nishimura, M., Hara-Nishimura, I., 2009. GNOM-LIKE1/ERMO1 and SEC24a/ERMO2 are required for maintenance of endoplasmic reticulum morphology in Arabidopsis thaliana. Plant Cell 21, 3672–3685. Nebenfu¨hr, A., Staehelin, L.A., 2001. Mobile factories: Golgi dynamics in plant cells. Trends Plant Sci. 6, 160–167. Nebenfu¨hr, A., Gallagher, L.A., Dunahay, T.G., Frohlick, J.A., Mazurkiewicz, A.M., Meehl, J.B., Staehelin, L.A., 1999. Stop-and-go movements of plant Golgi stacks are mediated by the acto-myosin system. Plant Physiol. 121, 1127–1141. Nebenfu¨hr, A., Frohlick, J.A., Staehelin, L.A., 2000. Redistribution of Golgi stacks and other organelles during mitosis and cytokinesis in plant cells. Plant Physiol. 124, 135–151. Nie, Z., Randazzo, P.A., 2006. Arf GAPs and membrane traffic. J. Cell Sci. 119, 1203–1211. Nilsson, T., Pypaert, M., Hoe, M.H., Slusarewicz, P., Berger, E.G., Warren, G., 1993. Overlapping distribution of two glycosyltransferases in the Golgi apparatus of HeLa cells. J. Cell Biol. 120, 5–13. Oikawa, A., Lund, C.H., Sakuragi, Y., Scheller, H.V., 2013. Golgi-localized enzyme complexes for plant cell wall biosynthesis. Trends Plant Sci. 18, 49–58. 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. Okamoto, M., Kurokawa, K., Matsuura-Tokita, K., Saito, C., Hirata, R., Nakano, A., 2012. High-curvature domains of the ER are important for the organization of ER exit sites in Saccharomyces cerevisiae. J. Cell Sci. 125, 3412–3420. Orci, L., Ravazzola, M., Meda, P., Holcomb, C., Moore, H.P., Hicke, L., Schekman, R., 1991. Mammalian Sec23p homologue is restricted to the endoplasmic reticulum transitional cytoplasm. Proc. Natl. Acad. Sci. U. S. A. 88, 8611–8615. Orci, L., Amherdt, M., Ravazzola, M., Perrelet, A., Rothman, J.E., 2000a. Exclusion of Golgi residents from transport vesicles budding from Golgi cisternae in intact cells. J. Cell Biol. 150, 1263–1270.

280

Yoko Ito et al.

Orci, L., Ravazzola, M., Volchuk, A., Engel, T., Gmachl, M., Amherdt, M., Perrelet, A., So¨llner, T.H., Rothman, J.E., 2000b. Anterograde flow of cargo across the Golgi stack potentially mediated via bidirectional “percolating” COPI vesicles. Proc. Natl. Acad. Sci. U. S. A. 97, 10400–10405. Osterrieder, A., Hummel, E., Carvalho, C.M., Hawes, C., 2010. Golgi membrane dynamics after induction of a dominant-negative mutant Sar1 GTPase in tobacco. J. Exp. Bot. 61, 405–422. Paccaud, J.P., Reith, W., Carpentier, J.L., Ravazzola, M., Amherdt, M., Schekman, R., Orci, L., 1996. Cloning and functional characterization of mammalian homologues of the COPII component Sec23. Mol. Cell. Biol. 7, 1535–1546. Paciorek, T., Zazimalova, E., Ruthardt, N., Petrasek, J., Stierhof, Y.-D., Kleine-Vehn, J., Morris, D.A., Emans, N., Jurgens, G., Geldner, N., Friml, J., 2005. Auxin inhibits endocytosis and promotes its own efflux from cells. Nature 435, 1251–1256. Pagano, A., Letourneur, F., Garcia-Estefania, D., Carpentier, J.-L., Orci, L., Paccaud, J.-P., 1999. Sec24 proteins and sorting at the endoplasmic reticulum. J. Biol. Chem. 274, 7833–7840. Palade, G., 1975. Intracellular aspects of the process of protein synthesis. Science 189, 347–358. Park, Y., Song, O.-K., Kwak, J., Hong, S., Lee, H., Nam, H., 1994. Functional complementation of a yeast vesicular transport mutation ypt1-1 by a Brassica napus cDNA clone encoding a small GTP-binding protein. Plant Mol. Biol. 26, 1725–1735. Park, M., Kim, S.J., Vitale, A., Hwang, I., 2004. Identification of the protein storage vacuole and protein targeting to the vacuole in leaf cells of three plant species. Plant Physiol. 134, 625–639. Parsons, H.T., Drakakaki, G., Heazlewood, J.L., 2013. Proteomic dissection of the Arabidopsis Golgi and trans-Golgi network. Front. Plant. Sci. 3, 298. Patterson, G.H., Hirschberg, K., Polishchuk, R.S., Gerlich, D., Phair, R.D., LippincottSchwartz, J., 2008. Transport through the Golgi apparatus by rapid partitioning within a two-phase membrane system. Cell 133, 1055–1067. Paulson, J.C., 2007. Innate immune response triggers lupus-like autoimmune disease. Cell 130, 589–591. Pavelka, M., Ellinger, A., Debbage, P., Loewe, C., Vetterlein, M., Roth, J., 1998. Endocytic routes to the Golgi apparatus. Histochem. Cell Biol. 109, 555–570. Pelletier, L., Stern, C.A., Pypaert, M., Sheff, D., Ngoˆ, H.M., Roper, N., He, C.Y., Hu, K., Toomre, D., Coppens, I., Roos, D.S., Joiner, K.A., Warren, G., 2002. Golgi biogenesis in Toxoplasma gondii. Nature 418, 548–552. Persico, A., Cervigni, R.I., Barretta, M.L., Colanzi, A., 2009. Mitotic inheritance of the Golgi complex. FEBS Lett. 583, 3857–3862. Pesacreta, T.C., Lucas, W.J., 1985. Presence of a partially-coated reticulum in angiosperms. Protoplasma 125, 173–184. Phillipson, B.A., Pimpl, P., daSilva, L.L.P., Crofts, A.J., Taylor, J.P., Movafeghi, A., Robinson, D.G., Denecke, J., 2001. Secretory bulk flow of soluble proteins is efficient and COPII dependent. Plant Cell 13, 2005–2020. Pimpl, P., Movafeghi, A., Coughlan, S., Denecke, J., Hillmer, S., Robinson, D.G., 2000. In situ localization and in vitro induction of plant COPI-coated vesicles. Plant Cell 12, 2219–2235. Pinheiro, H., Samalova, M., Geldner, N., Chory, J., Martinez, A., Moore, I., 2009. Genetic evidence that the higher plant Rab-D1 and Rab-D2 GTPases exhibit distinct but overlapping interactions in the early secretory pathway. J. Cell Sci. 122, 3749–3758. Polishchuk, R.S., Polishchuk, E.V., Mironov, A.A., 1999. Coalescence of Golgi fragments in microtubule-deprived living cells. Eur. J. Cell Biol. 78, 170–185. Popoff, V., Adolf, F., Bru¨gger, B., Wieland, F., 2011. COPI budding within the Golgi stack. Cold Spring Harb. Perspect. Biol. 3, a005231.

Golgi Apparatus of Plant Cells

281

Porter, K.R., 1961. The ground substance; observations from electron microscopy. In: Brachet, J., Mirsky, A.E. (Eds.), The Cell: Biochemistry, Physiology, Morphology, vol. II. Academic Press, New York, NY, pp. 621–675. Presley, J.F., Cole, N.B., Schroer, T.A., Hirschberg, K., Zaal, K.J.M., Lippincott-Schwartz, J., 1997. ER-to-Golgi transport visualized in living cells. Nature 389, 81–85. Preuss, D., Mulholland, J., Franzusoff, A., Segev, N., Botstein, D., 1992. Characterization of the Saccharomyces Golgi complex through the cell cycle by immunoelectron microscopy. Mol. Cell. Biol. 3, 789–803. Puri, S., Linstedt, A.D., 2003. Capacity of the Golgi apparatus for biogenesis from the endoplasmic reticulum. Mol. Cell. Biol. 14, 5011–5018. Puthenveedu, M.A., Bachert, C., Puri, S., Lanni, F., Linstedt, A.D., 2006. GM130 and GRASP65-dependent lateral cisternal fusion allows uniform Golgi-enzyme distribution. Nat. Cell Biol. 8, 238–248. Rabouille, C., Hui, N., Hunte, F., Kieckbusch, R., Berger, E.G., Warren, G., Nilsson, T., 1995. Mapping the distribution of Golgi enzymes involved in the construction of complex oligosaccharides. J. Cell Sci. 108, 1617–1627. Rambourg, A., Clermont, Y., Ke´pe`s, F., 1993. Modulation of the Golgi apparatus in Saccharomyces cerevisiae sec7 mutants as seen by three-dimensional electron microscopy. Anat. Rec. 237, 441–452. Reichardt, I., Stierhof, Y.-D., Mayer, U., Richter, S., Schwarz, H., Schumacher, K., Ju¨rgens, G., 2007. Plant cytokinesis requires de novo secretory trafficking but not endocytosis. Curr. Biol. 17, 2047–2053. Reilly, B.A., Kraynack, B.A., VanRheenen, S.M., Waters, M.G., 2001. Golgi-toendoplasmic reticulum (ER) retrograde traffic in yeast requires Dsl1p, a component of the ER target site that interacts with a COPI coat subunit. Mol. Cell. Biol. 12, 3783–3796. Reinhard, C., Schweikert, M., Wieland, F.T., Nickel, W., 2003. Functional reconstitution of COPI coat assembly and disassembly using chemically defined components. Proc. Natl. Acad. Sci. U. S. A. 100, 8253–8257. Ren, Y., Yip, C.K., Tripathi, A., Huie, D., Jeffrey, P.D., Walz, T., Hughson, F.M., 2009. A structure-based mechanism for vesicle capture by the multisubunit tethering complex Dsl1. Cell 139, 1119–1129. Renna, L., Hanton, S.L., Stefano, G., Bortolotti, L., Misra, V., Brandizzi, F., 2005. Identification and characterization of AtCASP, a plant transmembrane Golgi matrix protein. Plant Mol. Biol. 58, 109–122. Richter, S., Geldner, N., Schrader, J., Wolters, H., Stierhof, Y.D., Rios, G., Koncz, C., Robinson, D.G., Ju¨rgens, G., 2007. Functional diversification of closely related ARFGEFs in protein secretion and recycling. Nature 448, 488–492. Rios, R.M., Bornens, M., 2003. The Golgi apparatus at the cell centre. Curr. Opin. Cell Biol. 15, 60–66. Ritzenthaler, C., Nebenfu¨hr, A., Movafeghi, A., Stussi-Garaud, C., Behnia, L., Pimpl, P., Staehelin, L.A., Robinson, D.G., 2002. Reevaluation of the effects of brefeldin A on plant cells using tobacco Bright Yellow 2 cells expressing Golgi-targeted green fluorescent protein and COPI antisera. Plant Cell 14, 237–261. Rizzo, R., Parashuraman, S., Mirabelli, P., Puri, C., Lucocq, J., Luini, A., 2013. The dynamics of engineered resident proteins in the mammalian Golgi complex relies on cisternal maturation. J. Cell Biol. 201, 1027–1036. Roberg, K.J., Crotwell, M., Espenshade, P., Gimeno, R., Kaiser, C.A., 1999. LST1 is a SEC24 homologue used for selective export of the plasma membrane ATPase from the endoplasmic reticulum. J. Cell Biol. 145, 659–672. Robinson, D.G., Herranz, M.-C., Bubeck, J., Pepperkok, R., Ritzenthaler, C., 2007. Membrane dynamics in the early secretory pathway. Crit. Rev. Plant Sci. 26, 199–225.

282

Yoko Ito et al.

Robinson, D.G., Scheuring, D., Naramoto, S., Friml, J., 2011. ARF1 localizes to the Golgi and the trans-Golgi network. Plant Cell 23, 846. Rossanese, O.W., Soderholm, J., Bevis, B.J., Sears, I.B., O’Connor, J., Williamson, E.K., Glick, B.S., 1999. Golgi structure correlates with transitional endoplasmic reticulum organization in Pichia pastoris and Saccharomyces cerevisiae. J. Cell Biol. 145, 69–81. Roth, J., Taatjes, D.J., Lucocq, J.M., Weinstein, J., Paulson, J.C., 1985. Demonstration of an extensive trans-tubular network continuous with the Golgi apparatus stack that may function in glycosylation. Cell 43, 287–295. Rothman, J.E., 1994. Mechanisms of intracellular protein transport. Nature 372, 55–63. Sacher, M., Jiang, Y., Barrowman, J., Scarpa, A., Burston, J., Zhang, L., Schieltz, D., Yates, J.R., Abeliovich, H., Ferro-Novick, S., 1998. TRAPP, a highly conserved novel complex on the cis-Golgi that mediates vesicle docking and fusion. EMBO J. 17, 2494–2503. Saijo, Y., 2010. ER quality control of immune receptors and regulators in plants. Cell. Microbiol. 12, 716–724. Saint-Jore, C.M., Evins, J., Batoko, H., Brandizzi, F., Moore, I., Hawes, C., 2002. Redistribution of membrane proteins between the Golgi apparatus and endoplasmic reticulum in plants is reversible and not dependent on cytoskeletal networks. Plant J. 29, 661–678. Saint-Jore-Dupas, C., Nebenfu¨hr, A., Boulaflous, A., Follet-Gueye, M.L., Plasson, C., Hawes, C., Driouich, A., Faye, L., Gomord, V., 2006. Plant N-glycan processing enzymes employ different targeting mechanisms for their spatial arrangement along the secretory pathway. Plant Cell 18, 3182–3200. Saito, C., Ueda, T., 2009. Chapter 4: functions of RAB and SNARE proteins in plant life. Int. Rev. Cell Mol. Biol. 274, 183–233. Samalova, M., Fricker, M., Moore, I., 2006. Ratiometric fluorescence-imaging assays of plant membrane traffic using polyproteins. Traffic 7, 1701–1723. Samuelson, J., Banerjee, S., Magnelli, P., Cui, J., Kelleher, D.J., Gilmore, R., Robbins, P.W., 2005. The diversity of dolichol-linked precursors to Asn-linked glycans likely results from secondary loss of sets of glycosyltransferases. Proc. Natl. Acad. Sci. U. S. A. 102, 1548–1553. Sanderfoot, A.A., Assaad, F.F., Raikhel, N.V., 2000. The Arabidopsis genome. An abundance of soluble N-ethylmaleimide-sensitive factor adaptor protein receptors. Plant Physiol. 124, 1558–1569. Sapperstein, S.K., Lupashin, V.V., Schmitt, H.D., Waters, M.G., 1996. Assembly of the ER to Golgi SNARE complex requires Uso1p. J. Cell Biol. 132, 755–767. Saraste, J., Svensson, K., 1991. Distribution of the intermediate elements operating in ER to Golgi transport. J. Cell Sci. 100, 415–430. 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. Sato, K., Nishikawa, S., Nakano, A., 1995. Membrane protein retrieval from the Golgi apparatus to the endoplasmic reticulum (ER): characterization of the RER1 gene product as a component involved in ER localization of Sec12p. Mol. Cell. Biol. 6, 1459–1477. Sato, M., Sato, K., Nakano, A., 1996. Endoplasmic reticulum localization of Sec12p is achieved by two mechanisms: Rer1p-dependent retrieval that requires the transmembrane domain and Rer1p-independent retention that involves the cytoplasmic domain. J. Cell Biol. 134, 279–293. Sato, K., Ueda, T., Nakano, A., 1999. The Arabidopsis thaliana RER1 gene family: its potential role in the endoplasmic reticulum localization of membrane proteins. Plant Mol. Biol. 41, 815–824. Sato, K., Sato, M., Nakano, A., 2001. Rer1p, a retrieval receptor for endoplasmic reticulum membrane proteins. Is dynamically localized to the Golgi apparatus by coatomer. J. Cell Biol. 152, 935–944.

Golgi Apparatus of Plant Cells

283

Sato, K., Sato, M., Nakano, A., 2003. Rer1p, a retrieval receptor for ER membrane proteins, recognizes transmembrane domains in multiple modes. Mol. Cell. Biol. 14, 3605–3616. Satoh, A., Wang, Y., Malsam, J., Beard, M.B., Warren, G., 2003. Golgin-84 is a rab1 binding partner involved in Golgi sStructure. Traffic 4, 153–161. Scales, S.J., Pepperkok, R., Kreis, T.E., 1997. Visualization of ER-to-Golgi transport in living cells reveals a sequential mode of action for COPII and COPI. Cell 90, 1137–1148. Schekman, R., 1992. Genetic and biochemical analysis of vesicular traffic in yeast. Curr. Opin. Cell Biol. 4, 587–592. Schmitt, H.D., 2010. Dsl1p/Zw10: common mechanisms behind tethering vesicles and microtubules. Trends Cell Biol. 20, 257–268. Schmitz, K.R., Liu, J., Li, S., Setty, T.G., Wood, C.S., Burd, C.G., Ferguson, K.M., 2008. Golgi localization of glycosyltransferases requires a Vps74p oligomer. Dev. Cell 14, 523–534. Schoberer, J., Strasser, R., 2011. Sub-compartmental organization of Golgi-resident N-glycan processing enzymes in plants. Mol. Plant 4, 220. Schoberer, J., Vavra, U., Stadlmann, J., Hawes, C., Mach, L., Steinkellner, H., Strasser, R., 2009. Arginine/lysine residues in the cytoplasmic tail promote ER export of plant glycosylation enzymes. Traffic 10, 101–115. Schoberer, J., Runions, J., Steinkellner, H., Strasser, R., Hawes, C., Osterrieder, A., 2010. Sequential depletion and acquisition of proteins during Golgi stack disassembly and reformation. Traffic 11, 1429–1444. Schoberer, J., Liebminger, E., Botchway, S.W., Strasser, R., Hawes, C., 2013. Timeresolved fluorescence imaging reveals differential interactions of N-glycan processing enzymes across the Golgi stack in planta. Plant Physiol. 161, 1737–1754. Schweizer, A., Fransen, J.A., Matter, K., Kreis, T.E., Ginsel, L., Hauri, H.P., 1990. Identification of an intermediate compartment involved in protein transport from endoplasmic reticulum to Golgi apparatus. Eur. J. Cell Biol. 53, 185–196. Seemann, J., Pypaert, M., Taguchi, T., Malsam, J., Warren, G., 2002. Partitioning of the matrix fraction of the Golgi apparatus during mitosis in animal cells. Science 295, 848–851. Seguı´-Simarro, J.M., Staehelin, L.A., 2006. Cell cycle-dependent changes in Golgi stacks, vacuoles, clathrin-coated vesicles and multivesicular bodies in meristematic cells of Arabidopsis thaliana: a quantitative and spatial analysis. Planta 223, 223–236. Semenza, J.C., Hardwick, K.G., Dean, N., Pelham, H.R., 1990. ERD2, a yeast gene required for the receptor-mediated retrieval of luminal ER proteins from the secretory pathway. Cell 61, 1349–1357. Serafini, T., Stenbeck, G., Brecht, A., Lottspeich, F., Orel, L., Rothman, J.E., Wieland, F.T., 1991. A coat subunit of Golgi-derived non-clathrin-coated vesicles with homology to the clathrin-coated vesicle coat protein b-adaptin. Nature 349, 215–220. Shaywitz, D.A., Espenshade, P.J., Gimeno, R.E., Kaiser, C.A., 1997. COPII subunit interactions in the assembly of the vesicle coat. J. Biol. Chem. 272, 25413–25416. Shimmen, T., Yokota, E., 2004. Cytoplasmic streaming in plants. Curr. Opin. Cell Biol. 16, 68–72. Shorter, J., Warren, G., 1999. A role for the vesicle tethering protein, P115, in the postmitotic stacking of reassembling Golgi cisternae in a cell-free system. J. Cell Biol. 146, 57–70. Shorter, J., Watson, R., Giannakou, M.-E., Clarke, M., Warren, G., Barr, F.A., 1999. GRASP55, a second mammalian GRASP protein involved in the stacking of Golgi cisternae in a cell-free system. EMBO J. 18, 4949–4960. Shorter, J., Beard, M.B., Seemann, J., Dirac-Svejstrup, A.B., Warren, G., 2002. Sequential tethering of Golgins and catalysis of SNAREpin assembly by the vesicle-tethering protein p115. J. Cell Biol. 157, 45–62.

284

Yoko Ito et al.

Sjo¨strand, F.S., Hanzon, V., 1954. Ultrastructure of golgi apparatus of exocrinecells of mouse pancreas. Exp. Cell Res. 7, 415–429. Slusarewicz, P., Nilsson, T., Hui, N., Watson, R., Warren, G., 1994. Isolation of a matrix that binds medial Golgi enzymes. J. Cell Biol. 124, 405–413. Sohda, M., Misumi, Y., Yamamoto, A., Nakamura, N., Ogata, S., Sakisaka, S., Hirose, S., Ikehara, Y., Oda, K., 2010. Interaction of Golgin-84 with the COG complex mediates the intra-Golgi retrograde transport. Traffic 11, 1552–1566. Song, W., Henquet, M.G.L., Mentink, R.A., van Dijk, A.J., Cordewener, J.H.G., Bosch, D., America, A.H.P., van der Krol, A.R., 2011. N-glycoproteomics in plants: perspectives and challenges. J. Proteomics 74, 1463–1474. Sparkes, I., 2011. Recent advances in understanding plant myosin function: life in the fast lane. Mol. Plant 4, 805–812. Sparkes, I., Runions, J., Hawes, C., Griffing, L., 2009a. Movement and remodeling of the endoplasmic reticulum in nondividing cells of tobacco leaves. Plant Cell 21, 3937–3949. Sparkes, I.A., Ketelaar, T., De Ruijter, N.C.A., Hawes, C., 2009b. Grab a Golgi: laser trapping of Golgi bodies reveals in vivo interactions with the endoplasmic reticulum. Traffic 10, 567–571. Sparkes, I., Tolley, N., Aller, I., Svozil, J., Osterrieder, A., Botchway, S., Mueller, C., Frigerio, L., Hawes, C., 2010. Five Arabidopsis reticulon isoforms share endoplasmic reticulum location, topology, and membrane-shaping properties. Plant Cell 22, 1333–1343. Staehelin, L.A., Hepler, P.K., 1996. Cytokinesis in higher plants. Cell 84, 821–824. Staehelin, L.A., Kang, B.-H., 2008. Nanoscale Architecture of endoplasmic reticulum export sites and of Golgi membranes as determined by electron tomography. Plant Physiol. 147, 1454–1468. Staehelin, L.A., Giddings Jr., T.H., Kiss, J.Z., Sack, F.D., 1990. Macromolecular differentiation of Golgi stacks in root tips of Arabidopsis and Nicotiana seedlings as visualized in high pressure frozen and freeze-substituted samples. Protoplasma 157, 75–91. Stanley, H., Botas, J., Malhotra, V., 1997. The mechanism of Golgi segregation during mitosis is cell type-specific. Proc. Natl. Acad. Sci. U. S. A. 94, 14467–14470. Stefano, G., Renna, L., Chatre, L., Hanton, S.L., Moreau, P., Hawes, C., Brandizzi, F., 2006. In tobacco leaf epidermal cells, the integrity of protein export from the endoplasmic reticulum and of ER export sites depends on active COPI machinery. Plant J. 46, 95–110. Stierhof, Y.D., El Kasmi, F., 2010. Strategies to improve the antigenicity, ultrastructure preservation and visibility of trafficking compartments in Arabidopsis tissue. Eur. J. Cell Biol. 89, 285–297. Storrie, B., Micaroni, M., Morgan, Garry P., Jones, N., Kamykowski, Jeffrey A., Wilkins, N., Pan, T.H., Marsh, B.J., 2012. Electron tomography reveals Rab6 is essential to the trafficking of trans-Golgi clathrin and COPI-coated vesicles and the maintenance of Golgi cisternal number. Traffic 13, 727–744. Strasser, R., Mucha, J., Schwihla, H., Altmann, F., Glo¨ssl, J., Steinkellner, H., 1999. Molecular cloning and characterization of cDNA coding for b1,2N-acetylglucosaminyltransferase I (GlcNAc-TI) from Nicotiana tabacum. Glycobiology 9, 779–785. Strasser, R., Mucha, J., Mach, L., Altmann, F., Wilson, I.B.H., Glo¨ssl, J., Steinkellner, H., 2000. Molecular cloning and functional expression of b1,2-xylosyltransferase cDNA from Arabidopsis thaliana. FEBS Lett. 472, 105–108. Strasser, R., Altmann, F., Mach, L., Glo¨ssl, J., Steinkellner, H., 2004. Generation of Arabidopsis thaliana plants with complex N-glycans lacking b1,2-linked xylose and core a1,3-linked fucose. FEBS Lett. 561, 132–136. Strasser, R., Schoberer, J., Jin, C., Glo¨ssl, J., Mach, L., Steinkellner, H., 2006. Molecular cloning and characterization of Arabidopsis thaliana Golgi a-mannosidase II, a key enzyme in the formation of complex N-glycans in plants. Plant J. 45, 789–803.

Golgi Apparatus of Plant Cells

285

Strasser, R., Bondili, J.S., Vavra, U., Schoberer, J., Svoboda, B., Glo¨ssl, J., Le´onard, R., Stadlmann, J., Altmann, F., Steinkellner, H., Mach, L., 2007. A unique b1,3galactosyltransferase is indispensable for the biosynthesis of N-glycans containing lewis a structures in Arabidopsis thaliana. Plant Cell 19, 2278–2292. Strasser, R., Stadlmann, J., Scha¨hs, M., Stiegler, G., Quendler, H., Mach, L., Glo¨ssl, J., Weterings, K., Pabst, M., Steinkellner, H., 2008. Generation of glyco-engineered Nicotiana benthamiana for the production of monoclonal antibodies with a homogeneous human-like N-glycan structure. Plant Biotechnol. J. 6, 392–402. Strasser, R., Castilho, A., Stadlmann, J., Kunert, R., Quendler, H., Gattinger, P., Jez, J., Rademacher, T., Altmann, F., Mach, L., Steinkellner, H., 2009. Improved virus neutralization by plant-produced anti-HIV antibodies with a homogeneous b1,4galactosylated N-glycan profile. J. Biol. Chem. 284, 20479–20485. Su¨tterlin, C., Colanzi, A., 2010. The Golgi and the centrosome: building a functional partnership. J. Cell Biol. 188, 621–628. Su¨tterlin, C., Polishchuk, R., Pecot, M., Malhotra, V., 2005. The Golgi-associated protein GRASP65 regulates spindle dynamics and is essential for cell division. Mol. Cell. Biol. 16, 3211–3222. Suvorova, E.S., Duden, R., Lupashin, V.V., 2002. The Sec34/Sec35p complex, a Ypt1p effector required for retrograde intra-Golgi trafficking, interacts with Golgi SNAREs and COPI vesicle coat proteins. J. Cell Biol. 157, 631–643. Sztul, E., Lupashin, V., 2009. Role of vesicle tethering factors in the ER–Golgi membrane traffic. FEBS Lett. 583, 3770–3783. Szul, T., Sztul, E., 2011. COPII and COPI Traffic at the ER-Golgi Interface. Physiology 26, 348–364. Taatjes, D.J., Roth, J., 1986. The trans-tubular network of the hepatocyte Golgi apparatus is part of the secretory pathway. Eur. J. Cell Biol. 42, 344–350. Tai, W.C.S., Banfield, D.K., 2001. AtBS14a and AtBS14b, two Bet1/Sft1-like SNAREs from Arabidopsis thaliana that complement mutations in the yeast SFT1 gene. FEBS Lett. 500, 177–182. Takagi, J., Renna, L., Takahashi, H., Koumoto, Y., Tamura, K., Stefano, G., Fukao, Y., Kondo, M., Nishimura, M., Shimada, T., Brandizzi, F., Hara-Nishimura, I., 2013. MAIGO5 functions in protein export from Golgi-associated endoplasmic reticulum exit sites in Arabidopsis. Plant Cell 25, 4658–4675. Takahashi, H., Tamura, K., Takagi, J., Koumoto, Y., Hara-Nishimura, I., Shimada, T., 2010. MAG4/Atp115 is a golgi-localized tethering factor that mediates efficient anterograde transport in Arabidopsis. Plant Cell Physiol. 51, 1777–1787. Takeuchi, M., Tada, M., Saito, C., Yashiroda, H., Nakano, A., 1998. Isolation of a tobacco cDNA encoding Sar1 GTPase and analysis of its dominant mutations in vesicular traffic using a yeast complementation system. Plant Cell Physiol. 39, 590–599. Takeuchi, M., Ueda, T., Sato, K., Abe, H., Nagata, T., Nakano, A., 2000. A dominant negative mutant of Sar1 GTPase inhibits protein transport from the endoplasmic reticulum to the Golgi apparatus in tobacco and Arabidopsis cultured cells. Plant J. 23, 517–525. 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. Tanaka, H., Kitakura, S., De Rycke, R., De Groodt, R., Friml, J., 2009. Fluorescence imaging-based screen identifies ARF GEF component of early endosomal trafficking. Curr. Biol. 19, 391–397. Tanaka, Y., Nishimura, K., Kawamukai, M., Oshima, A., Nakagawa, T., 2013. Redundant function of two Arabidopsis COPII components, AtSec24B and AtSec24C, is essential for male and female gametogenesis. Planta 238, 561–575.

286

Yoko Ito et al.

Tang, B.L., Wong, S.H., Qi, X.L., Low, S.H., Hong, W., 1993. Molecular cloning, characterization, subcellular localization and dynamics of p23, the mammalian KDEL receptor. J. Cell Biol. 120, 325–338. Teh, O., Moore, I., 2007. An ARF-GEF acting at the Golgi and in selective endocytosis in polarized plant cells. Nature 448, 493–496. Tian, L., Dai, L.L., Yin, Z.J., Fukuda, M., Kumamaru, T., Dong, X.B., Xu, X.P., Qu, L.Q., 2013. Small GTPase Sar1 is crucial for proglutelin and a-globulin export from the endoplasmic reticulum in rice endosperm. J. Exp. Bot. 64, 2831–2845. Tominaga, M., Nakano, A., 2012. Plant-specific myosin XI, a molecular perspective. Front. Plant. Sci. 3, 211. Tu, L., Banfield, D., 2010. Localization of Golgi-resident glycosyltransferases. Cell. Mol. Life Sci. 67, 29–41. Tu, L., Tai, W.C.S., Chen, L., Banfield, D.K., 2008. Signal-mediated dynamic retention of glycosyltransferases in the Golgi. Science 321, 404–407. Ueda, H., Yokota, E., Kutsuna, N., Shimada, T., Tamura, K., Shimmen, T., Hasezawa, S., Dolja, V.V., Hara-Nishimura, I., 2010. Myosin-dependent endoplasmic reticulum motility and F-actin organization in plant cells. Proc. Natl. Acad. Sci. U. S. A. 107, 6894–6899. Uemura, T., Nakano, A., 2013. Plant TGNs: dynamics and physiological functions. Histochem. Cell Biol. 140, 341–345. Uemura, T., Ueda, T., Ohniwa, R.L., Nakano, A., Takeyasu, K., Sato, M.H., 2004. Systematic analysis of SNARE molecules in Arabidopsis: dissection of the post-Golgi network in plant cells. Cell Struct. Funct. 29, 49–65. Uemura, T., Suda, Y., Ueda, T., Nakano, A. Dynamic behavior of the trans-Golgi network in root tissues of Arabidopsis revealed by super-resolution live imaging. Plant Cell Physiol. in press. Vasile, E., Oka, T., Ericsson, M., Nakamura, N., Krieger, M., 2006. IntraGolgi distribution of the conserved oligomeric Golgi (COG) complex. Exp. Cell Res. 312, 3132–3141. Vinke, F., Grieve, A., Rabouille, C., 2011. The multiple facets of the Golgi reassembly stacking proteins. Biochem. J. 433, 423–433. Viotti, C., Bubeck, J., Stierhof, Y.D., Krebs, M., Langhans, M., van den Berg, W., van Dongen, W., Richter, S., Geldner, N., Takano, J., Jurgens, G., de Vries, S.C., Robinson, D.G., Schumacher, K., 2010. Endocytic and secretory traffic in Arabidopsis merge in the trans-Golgi network/early endosome, an independent and highly dynamic organelle. Plant Cell 22, 1344–1357. von Schaewen, A., Sturm, A., O’Neill, J., Chrispeels, M.J., 1993. Isolation of a mutant Arabidopsis plant that lacks N-acetyl glucosaminyl transferase I and is unable to synthesize golgi-modified complex N-linked glycans. Plant Physiol. 102, 1109–1118. Wang, W., Sacher, M., Ferro-Novick, S., 2000. Trapp stimulates guanine nucleotide exchange on Ypt1p. J. Cell Biol. 151, 289–296. Wang, Y., Seemann, J., Pypaert, M., Shorter, J., Warren, G., 2003. A direct role for GRASP65 as a mitotically regulated Golgi stacking factor. EMBO J. 22, 3279–3290. Wang, Y., Wei, J.-H., Bisel, B., Tang, D., Seemann, J., 2008. Golgi cisternal unstacking stimulates COPI vesicle budding and protein transport. PLoS ONE 3, e1647. Waters, M.G., Serafini, T., Rothman, J.E., 1991. ’Coatomer’: a cytosolic protein complex containing subunits of non-clathrin-coated Golgi transport vesicles. Nature 349, 248–251. Watson, P., Townley, A.K., Koka, P., Palmer, K.J., Stephens, D.J., 2006. Sec16 defines endoplasmic reticulum exit sites and is required for secretory cargo export in mammalian cells. Traffic 7, 1678–1687. Wee, E.G.-T., Sherrier, D.J., Prime, T.A., Dupree, P., 1998. Targeting of active sialyltransferase to the plant Golgi apparatus. Plant Cell 10, 1759–1768. Wei, J.-H., Seemann, J., 2009. The mitotic spindle mediates inheritance of the Golgi ribbon structure. J. Cell Biol. 184, 391–397.

Golgi Apparatus of Plant Cells

287

Wei, J.-H., Seemann, J., 2010. Unraveling the Golgi ribbon. Traffic 11, 1391–1400. Wendeler, M.W., Paccaud, J.P., Hauri, H.P., 2007. Role of Sec24 isoforms in selective export of membrane proteins from the endoplasmic reticulum. EMBO Rep. 8, 258–264. Witte, K., Schuh, A.L., Hegermann, J., Sarkeshik, A., Mayers, J.R., Schwarze, K., Yates III, J.R., Eimer, S., Audhya, A., 2011. TFG-1 function in protein secretion and oncogenesis. Nat. Cell Biol. 13, 550–558. Wood, S.A., Park, J.E., Brown, W.J., 1991. Brefeldin A causes a microtubule-mediated fusion of the trans-Golgi network and early endosomes. Cell 67, 591–600. Xiang, Y., Wang, Y., 2010. GRASP55 and GRASP65 play complementary and essential roles in Golgi cisternal stacking. J. Cell Biol. 188, 237–251. Xiang, Y., Zhang, X., Nix, D.B., Katoh, T., Aoki, K., Tiemeyer, M., Wang, Y., 2013. Regulation of protein glycosylation and sorting by the Golgi matrix proteins GRASP55/65. Nat. Commun. 4, 1659. Xu, D., Hay, J.C., 2004. Reconstitution of COPII vesicle fusion to generate a pre-Golgi intermediate compartment. J. Cell Biol. 167, 997–1003. Xu, J., Scheres, B., 2005. Dissection of Arabidopsis ADP-ribosylation factor 1 function in epidermal cell polarity. Plant Cell 17, 525–536. Xu, Y., Martin, S., James, D.E., Hong, W., 2002. GS15 forms a SNARE complex with syntaxin 5, GS28, and Ykt6 and is implicated in traffic in the early cisternae of the Golgi apparatus. Mol. Cell. Biol. 13, 3493–3507. Yadav, S., Puri, S., Linstedt, A.D., 2009. A primary role for Golgi positioning in directed secretion, cell polarity, and wound healing. Mol. Cell. Biol. 20, 1728–1736. Yamasaki, A., Menon, S., Yu, S., Barrowman, J., Meerloo, T., Oorschot, V., Klumperman, J., Satoh, A., Ferro-Novick, S., 2009. mTrs130 is a component of a mammalian TRAPPII complex, a Rab1 GEF that binds to COPI-coated vesicles. Mol. Cell. Biol. 20, 4205–4215. Yang, Y.D., Elamawi, R., Bubeck, J., Pepperkok, R., Ritzenthaler, C., Robinson, D.G., 2005. Dynamics of COPII vesicles and the Golgi apparatus in cultured Nicotiana tabacum BY-2 cells provides evidence for transient association of Golgi stacks with endoplasmic reticulum exit sites. Plant Cell 17, 1513–1531. Yano, H., Yamamoto-Hino, M., Abe, M., Kuwahara, R., Haraguchi, S., Kusaka, I., Awano, W., Kinoshita-Toyoda, A., Toyoda, H., Goto, S., 2005. Distinct functional units of the Golgi complex in Drosophila cells. Proc. Natl. Acad. Sci. U. S. A. 102, 13467–13472. Yorimitsu, T., Sato, K., 2012. Insights into structural and regulatory roles of Sec16 in COPII vesicle formation at ER exit sites. Mol. Cell. Biol. 23, 2930–2942. Yoshihisa, T., Barlowe, C., Schekman, R., 1993. Requirement for a GTPase-activating protein in vesicle budding from the endoplasmic reticulum. Science 259, 1466–1468. Yuasa, K., Toyooka, K., Fukuda, H., Matsuoka, K., 2005. Membrane-anchored prolyl hydroxylase with an export signal from the endoplasmic reticulum. Plant J. 41, 81–94. Zaal, K.J.M., Smith, C.L., Polishchuk, R.S., Altan, N., Cole, N.B., Ellenberg, J., Hirschberg, K., Presley, J.F., Roberts, T.H., Siggia, E., Phair, R.D., LippincottSchwartz, J., 1999. Golgi membranes are absorbed into and reemerge from the ER during mitosis. Cell 99, 589–601. Zhang, G.F., Staehelin, L.A., 1992. Functional compartmentation of the Golgi apparatus of plant cells: immunocytochemical analysis of high-pressure frozen- and freeze-substituted sycamore maple suspension culture cells. Plant Physiol. 99, 1070–1083. Zheng, H., Camacho, L., Wee, E., Batoko, H., Legen, J., Leaver, C.J., Malho´, R., Hussey, P.J., Moore, I., 2005. A Rab-E GTPase mutant acts downstream of the Rab-D subclass in biosynthetic membrane traffic to the plasma membrane in tobacco leaf epidermis. Plant Cell 17, 2020–2036. Zolov, S.N., Lupashin, V.V., 2005. Cog3p depletion blocks vesicle-mediated Golgi retrograde trafficking in HeLa cells. J. Cell Biol. 168, 747–759.

Formation and maintenance of the Golgi apparatus in plant cells.

The Golgi apparatus plays essential roles in intracellular trafficking, protein and lipid modification, and polysaccharide synthesis in eukaryotic cel...
2MB Sizes 3 Downloads 4 Views