首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
2.
3.
The mitochondrial inner membrane contains a large protein complex that functions in inner membrane organization and formation of membrane contact sites. The complex was variably named the mitochondrial contact site complex, mitochondrial inner membrane organizing system, mitochondrial organizing structure, or Mitofilin/Fcj1 complex. To facilitate future studies, we propose to unify the nomenclature and term the complex “mitochondrial contact site and cristae organizing system” and its subunits Mic10 to Mic60.Mitochondria possess two membranes of different architecture and function (Palade, 1952; Hackenbrock, 1968). Both membranes work together for essential shared functions, such as protein import (Schatz, 1996; Neupert and Herrmann, 2007; Chacinska et al., 2009). The outer membrane harbors machinery that controls the shape of the organelle and is crucial for the communication of mitochondria with the rest of the cell. The inner membrane harbors the complexes of the respiratory chain, the F1Fo-ATP synthase, numerous metabolite carriers, and enzymes of mitochondrial metabolism. It consists of two domains: the inner boundary membrane, which is adjacent to the outer membrane, and invaginations of different shape, termed cristae (Werner and Neupert, 1972; Frey and Mannella, 2000; Hoppins et al., 2007; Pellegrini and Scorrano, 2007; Zick et al., 2009; Davies et al., 2011). Tubular openings, termed crista junctions (Perkins et al., 1997), connect inner boundary membrane and cristae membranes (Fig. 1, A and B). Respiratory chain complexes and the F1Fo-ATP synthase are preferentially located in the cristae membranes, whereas preprotein translocases are enriched in the inner boundary membrane (Vogel et al., 2006; Wurm and Jakobs, 2006; Davies et al., 2011). Contact sites between outer membrane and inner boundary membrane promote import of preproteins, metabolite channeling, lipid transport, and membrane dynamics (Frey and Mannella, 2000; Sesaki and Jensen, 2004; Hoppins et al., 2007, 2011; Neupert and Herrmann, 2007; Chacinska et al., 2009; Connerth et al., 2012; van der Laan et al., 2012).Open in a separate windowFigure 1.MICOS complex. (A) The MICOS complex (hypothetical model), previously also termed MINOS, MitOS, or Mitofilin/Fcj1 complex, is required for maintenance of the characteristic architecture of the mitochondrial inner membrane (IM) and forms contact sites with the outer membrane (OM). In budding yeast, six subunits of MICOS have been identified. All subunits are exposed to the intermembrane space (IMS), five are integral inner membrane proteins (Mic10, Mic12, Mic26, Mic27, and Mic60), and one is a peripheral inner membrane protein (Mic19). Mic26 is related to Mic27; however, mic26Δ yeast cells show considerably less severe defects of mitochondrial inner membrane architecture than mic27Δ cells (Harner et al., 2011; Hoppins et al., 2011; von der Malsburg et al., 2011). The MICOS complex of metazoa additionally contains Mic25, which is related to Mic19, yet subunits corresponding to Mic12 and Mic26 have not been identified so far. MICOS subunits that have been conserved in most organisms analyzed are indicated by bold boundary lines. (B, top) Wild-type architecture of the mitochondrial inner membrane with crista junctions and cristae. (bottom) This architecture is considerably altered in micos mutant mitochondria: most cristae membranes are detached from the inner boundary membrane and form internal membrane stacks. In some micos mutants (deficiency of mammalian Mic19 or Mic25), a loss of cristae membranes was observed (Darshi et al., 2011; An et al., 2012). Figure by M. Bohnert (Institute of Biochemistry and Molecular Biology, University of Freiburg, Freiburg, Germany).To understand the complex architecture of mitochondria, it will be crucial to identify the molecular machineries that control the interaction between mitochondrial outer and inner membranes and the characteristic organization of the inner membrane. A convergence of independent studies led to the identification of a large heterooligomeric protein complex of the mitochondrial inner membrane conserved from yeast to humans that plays crucial roles in the maintenance of crista junctions, inner membrane architecture, and formation of contact sites to the outer membrane (Fig. 1 A). Several names were used by different research groups to describe the complex, including mitochondrial contact site (MICOS) complex, mitochondrial inner membrane organizing system (MINOS), mitochondrial organizing structure (MitOS), Mitofilin complex, or Fcj1 (formation of crista junction protein 1) complex (Harner et al., 2011; Hoppins et al., 2011; von der Malsburg et al., 2011; Alkhaja et al., 2012). Mitofilin, also termed Fcj1, was the first component identified (Icho et al., 1994; Odgren et al., 1996; Gieffers et al., 1997; John et al., 2005) and was observed enriched at crista junctions (Rabl et al., 2009). Mutants of Mitofilin/Fcj1 as well as of other MICOS/MINOS/MitOS subunits show a strikingly altered inner membrane architecture. They lose crista junctions and contain large internal membrane stacks, the respiratory activity is reduced, and mitochondrial DNA nucleoids are altered (Fig. 1 B; John et al., 2005; Hess et al., 2009; Rabl et al., 2009; Mun et al., 2010; Harner et al., 2011; Head et al., 2011; Hoppins et al., 2011; von der Malsburg et al., 2011; Alkhaja et al., 2012; Itoh et al., 2013). It has been reported that the complex interacts with a variety of outer membrane proteins, such as channel proteins and components of the protein translocases and mitochondrial fusion machines, and defects impair the biogenesis of mitochondrial proteins (Xie et al., 2007; Darshi et al., 2011; Harner et al., 2011; Hoppins et al., 2011; von der Malsburg et al., 2011; Alkhaja et al., 2012; An et al., 2012; Bohnert et al., 2012; Körner et al., 2012; Ott et al., 2012; Zerbes et al., 2012; Jans et al., 2013; Weber et al., 2013). The MICOS/MINOS/MitOS/Mitofilin/Fcj1 complex thus plays crucial roles in mitochondrial architecture, dynamics, and biogenesis. However, communication of results in this rapidly developing field has been complicated by several different nomenclatures used for the complex as well as for its subunits (
Standard nameFormer namesYeast ORFReferences
Complex
MICOSMINOS, MitOS, MIB, Mitofilin complex, and Fcj1 complexXie et al., 2007; Rabl et al., 2009; Darshi et al., 2011; Harner et al., 2011; Hoppins et al., 2011; von der Malsburg et al., 2011; Alkhaja et al., 2012; An et al., 2012; Bohnert et al., 2012; Ott et al., 2012; Jans et al., 2013; Weber et al., 2013
Subunits
Mic10Mcs10, Mio10, Mos1, and MINOS1YCL057C-AHarner et al., 2011; Hoppins et al., 2011; von der Malsburg et al., 2011; Alkhaja et al., 2012; Itoh et al., 2013; Jans et al., 2013; Varabyova et al., 2013
Mic12Aim5, Fmp51, and Mcs12YBR262CHess et al., 2009; Harner et al., 2011; Hoppins et al., 2011; von der Malsburg et al., 2011; Varabyova et al., 2013
Mic19Aim13, Mcs19, CHCH-3, CHCHD3, and MINOS3YFR011CXie et al., 2007; Hess et al., 2009; Darshi et al., 2011; Head et al., 2011; Alkhaja et al., 2012; Ott et al., 2012; Jans et al., 2013; Varabyova et al., 2013
Mic25 (metazoan Mic19 homologue)CHCHD6 and CHCM1Xie et al., 2007; An et al., 2012
Mic26Mcs29, Mio27, and Mos2YGR235CHarner et al., 2011; Hoppins et al., 2011; von der Malsburg et al., 2011
Mic27Aim37, Mcs27, APOOL, and MOMA-1YNL100WHess et al., 2009; Harner et al., 2011; Head et al., 2011; Hoppins et al., 2011; von der Malsburg et al., 2011; Weber et al., 2013
Mic60Fcj1, Aim28, Fmp13, Mitofilin, HMP, IMMT, and MINOS2YKR016WIcho et al., 1994; Odgren et al., 1996; Gieffers et al., 1997; John et al., 2005; Wang et al., 2008; Rabl et al., 2009; Rossi et al., 2009; Mun et al., 2010; Park et al., 2010; Körner et al., 2012; Zerbes et al., 2012; Itoh et al., 2013; Varabyova et al., 2013
Open in a separate windowAPOOL, apolipoprotein O–like; HMP, heart muscle protein; IMMT, inner mitochondrial membrane protein; MIB, mitochondrial intermembrane space bridging.To rectify this situation, all authors of this article have agreed on a new uniform nomenclature with the following guidelines. (a) The complex will be called “mitochondrial contact site and cristae organizing system” (MICOS). The protein subunits of MICOS are named Mic10 to Mic60 as listed in Gabriel et al., 2007; Vögtle et al., 2012) will be changed to Mix14, Mix17, and Mix23 (mitochondrial intermembrane space CXnC motif proteins) in the Saccharomyces Genome Database, and the new nomenclature will be used for orthologues identified in other organisms.The MICOS complex is of central importance for the maintenance of mitochondrial inner membrane architecture and the formation of contact sites between outer and inner membranes and thus is involved in the regulation of mitochondrial dynamics, biogenesis, and inheritance. We expect that the uniform nomenclature will facilitate future studies on mitochondrial membrane architecture and dynamics.  相似文献   

4.
SOS2-LIKE PROTEIN KINASE5, an SNF1-RELATED PROTEIN KINASE3-Type Protein Kinase,Is Important for Abscisic Acid Responses in Arabidopsis through Phosphorylation of ABSCISIC ACID-INSENSITIVE5   总被引:1,自引:0,他引:1  
Xiaona Zhou  Hongmei Hao  Yuguo Zhang  Yili Bai  Wenbo Zhu  Yunxia Qin  Feifei Yuan  Feiyi Zhao  Mengyao Wang  Jingjiang Hu  Hong Xu  Aiguang Guo  Huixian Zhao  Yang Zhao  Cuiling Cao  Yongqing Yang  Karen S. Schumaker  Yan Guo  Chang Gen Xie 《Plant physiology》2015,168(2):659-676
  相似文献   

5.
Evolution and Function of the Plant Cell Wall Synthesis-Related Glycosyltransferase Family 8     
Yanbin Yin  Huiling Chen  Michael G. Hahn  Debra Mohnen  Ying Xu 《Plant physiology》2010,153(4):1729-1746
Carbohydrate-active enzyme glycosyltransferase family 8 (GT8) includes the plant galacturonosyltransferase1-related gene family of proven and putative α-galacturonosyltransferase (GAUT) and GAUT-like (GATL) genes. We computationally identified and investigated this family in 15 fully sequenced plant and green algal genomes and in the National Center for Biotechnology Information nonredundant protein database to determine the phylogenetic relatedness of the GAUTs and GATLs to other GT8 family members. The GT8 proteins fall into three well-delineated major classes. In addition to GAUTs and GATLs, known or predicted to be involved in plant cell wall biosynthesis, class I also includes a lower plant-specific GAUT and GATL-related (GATR) subfamily, two metazoan subfamilies, and proteins from other eukaryotes and cyanobacteria. Class II includes galactinol synthases and plant glycogenin-like starch initiation proteins that are not known to be directly involved in cell wall synthesis, as well as proteins from fungi, metazoans, viruses, and bacteria. Class III consists almost entirely of bacterial proteins that are lipooligo/polysaccharide α-galactosyltransferases and α-glucosyltransferases. Sequence motifs conserved across all GT8 subfamilies and those specific to plant cell wall-related GT8 subfamilies were identified and mapped onto a predicted GAUT1 protein structure. The tertiary structure prediction identified sequence motifs likely to represent key amino acids involved in catalysis, substrate binding, protein-protein interactions, and structural elements required for GAUT1 function. The results show that the GAUTs, GATLs, and GATRs have a different evolutionary origin than other plant GT8 genes, were likely acquired from an ancient cyanobacterium (Synechococcus) progenitor, and separate into unique subclades that may indicate functional specialization.Plant cell walls are composed of three principal types of polysaccharides: cellulose, hemicellulose, and pectin. Studying the biosynthesis and degradation of these biopolymers is important because cell walls have multiple roles in plants, including providing structural support to cells and defense against pathogens, serving as cell-specific developmental and differentiation markers, and mediating or facilitating cell-cell communication. In addition to their important roles within plants, cell walls also have many economic uses in human and animal nutrition and as sources of natural textile fibers, paper and wood products, and components of fine chemicals and medicinal products. The study of the biosynthesis and biodegradation of plant cell walls has become even more significant because cell walls are the major components of biomass (Mohnen et al., 2008), which is the most promising renewable source for the production of biofuels and biomaterials (Ragauskas et al., 2006; Pauly and Keegstra, 2008). Analyses of fully sequenced plant genomes have revealed that they encode hundreds or even thousands of carbohydrate-active enzymes (CAZy; Henrissat et al., 2001; Yokoyama and Nishitani, 2004; Geisler-Lee et al., 2006). Most of these CAZy enzymes (Cantarel et al., 2009) are glycosyltransferases (GTs) or glycoside hydrolases, which are key players in plant cell wall biosynthesis and modification (Cosgrove, 2005).The CAZy database is classified into 290 protein families (www.cazy.org; release of September 2008), of which 92 are GT families (Cantarel et al., 2009). A number of the GT families have been previously characterized to be involved in plant cell wall biosynthesis. For example, the GT2 family is known to include cellulose synthases and some hemicellulose backbone synthases (Lerouxel et al., 2006), such as mannan synthases (Dhugga et al., 2004; Liepman et al., 2005), putative xyloglucan synthases (Cocuron et al., 2007), and mixed linkage glucan synthases (Burton et al., 2006). With respect to the synthesis of xylan, a type of hemicellulose, four Arabidopsis (Arabidopsis thaliana) proteins from the GT43 family, irregular xylem 9 (IRX9), IRX14, IRX9-L, and IRX14-L, and two proteins from the GT47 family, IRX10 and IRX10-L, are candidates (York and O''Neill, 2008) for glucuronoxylan backbone synthases (Brown et al., 2007, 2009; Lee et al., 2007a; Peña et al., 2007; Wu et al., 2009). In addition, three proteins have been implicated in the synthesis of an oligosaccharide thought to act either as a primer or terminator in xylan synthesis (Peña et al., 2007): two from the GT8 family (IRX8/GAUT12 [Persson et al., 2007] and PARVUS/GATL1 [Brown et al., 2007; Lee et al., 2007b]) and one from the GT47 family (FRA8/IRX7 [Zhong et al., 2005]).The GT families involved in the biosynthesis of pectins have been relatively less studied until recently. In 2006, a gene in CAZy family GT8 was shown to encode a functional homogalacturonan α-galacturonosyltransferase, GAUT1 (Sterling et al., 2006). GAUT1 belongs to a 25-member gene family in Arabidopsis, the GAUT1-related gene family, that includes two distinct but closely related families, the galacturonosyltransferase (GAUT) genes and the galacturonosyltransferase-like (GATL) genes (Sterling et al., 2006). Another GAUT gene, GAUT8/QUA1, has been suggested to be involved in pectin and/or xylan synthesis, based on the phenotypes of plant lines carrying mutations in this gene (Bouton et al., 2002; Orfila et al., 2005). It has further been suggested that multiple members of the GT8 family are galacturonosyltransferases involved in pectin and/or xylan biosynthesis (Mohnen, 2008; Caffall and Mohnen, 2009; Caffall et al., 2009).Aside from the 25 GAUT and GATL genes, Arabidopsis has 16 other family GT8 genes, according to the CAZy database, which do not seem to have the conserved sequence motifs found in GAUTs and GATLs: HxxGxxKPW and GLG (Sterling et al., 2006). Eight of these 16 genes are annotated as galactinol synthase (GolS) by The Arabidopsis Information Resource (TAIR; www.arabidopsis.org), and three of these AtGolS enzymes have been implicated in the synthesis of raffinose family oligosaccharides that are associated with stress tolerance (Taji et al., 2002). The other eight Arabidopsis GT8 genes are annotated as plant glycogenin-like starch initiation proteins (PGSIPs) in TAIR. PGSIPs have been proposed to be involved in the synthesis of primers necessary for starch biosynthesis (Chatterjee et al., 2005). Hence, the GT8 family is a protein family consisting of enzymes with very distinct proven and proposed functions. Indeed, a suggestion has been made to split the GT8 family into two groups (Sterling et al., 2006), namely, the cell wall biosynthesis-related genes (GAUTs and GATLs) and the non-cell wall synthesis-related genes (GolSs and PGSIPs).We are interested in further defining the functions of the GAUT and GATL proteins in plants, in particular their role(s) in plant cell wall synthesis. The apparent disparate functions of the GT8 family (i.e. the GAUTs and GATLs as proven and putative plant cell wall polysaccharide biosynthetic α-galacturonosyltransferases, the eukaryotic GolSs as α-galactosyltransferases that synthesize the first step in the synthesis of the oligosaccharides stachyose and raffinose, the putative PGSIPs, and the large bacterial GT8 family of diverse α-glucosyltransferases and α-galactosyltransferases involved in lipopolysaccharide and lipooligosaccharide synthesis) indicate that the GT8 family members are involved in several unique types of glycoconjugate and glycan biosynthetic processes (Yin et al., 2010). This observation led us to ask whether any of the GT8 family members are sufficiently closely related to GAUT and GATL genes to be informative regarding GAUT or GATL biosynthetic function(s) and/or mechanism(s).To investigate the relatedness of the members of the GT8 gene family, we carried out a detailed phylogenetic analysis of the entire GT8 family in 15 completely sequenced plant and green algal genomes (AbbreviationCladeSpeciesGenome PublishedDownloaded frommpcGreen algaeMicromonas pusilla CCMP1545Worden et al. (2009)JGI version 2.0mprGreen algaeMicromonas strain RCC299Worden et al. (2009)JGI version 2.0olGreen algaeOstreococcus lucimarinusPalenik et al. (2007)JGI version 1.0otGreen algaeOstreococcus tauriDerelle et al. (2006)JGI version 1.0crGreen algaeChlamydomonas reinhardtiiMerchant et al. (2007)JGI version 3.0vcGreen algaeVolvox carteri f. nagariensisNoJGI version 1.0ppMossPhyscomitrella patens ssp. patensRensing et al. (2008)JGI version 1.1smSpike mossSelaginella moellendorffiiNoJGI version 1.0ptDicotPopulus trichocarpaTuskan et al. (2006)JGI version 1.1atDicotArabidopsis thalianaArabidopsis Genome Initiative (2000)TAIR version 9.0vvDicotVitis viniferaJaillon et al. (2007)http://www.genoscope.cns.fr/gmDicotGlycine maxSchmutz et al. (2010)JGI version 1.0osMonocotOryza sativaGoff et al. (2002); Yu et al. (2002)TIGR version 6.1sbMonocotSorghum bicolorPaterson et al. (2009)JGI version 1.0bdMonocotBrachypodium distachyonVogel et al. (2010)JGI version 1.0Open in a separate window  相似文献   

6.
Peaks cloaked in the mist: The landscape of mammalian replication origins     
Olivier Hyrien 《The Journal of cell biology》2015,208(2):147-160
  相似文献   

7.
Cell biology in neuroscience: RNA-based mechanisms underlying axon guidance     
Toshiaki Shigeoka  Bo Lu  Christine E. Holt 《The Journal of cell biology》2013,202(7):991-999
  相似文献   

8.
The cell biology of disease: The cellular and molecular basis for malaria parasite invasion of the human red blood cell     
Alan F. Cowman  Drew Berry  Jake Baum 《The Journal of cell biology》2012,198(6):961-971
  相似文献   

9.
The cell biology of disease: Lysosomal storage disorders: The cellular impact of lysosomal dysfunction     
Frances M. Platt  Barry Boland  Aarnoud C. van der Spoel 《The Journal of cell biology》2012,199(5):723-734
  相似文献   

10.
Mitotic spindle (DIS)orientation and DISease: Cause or consequence?     
Anna Noatynska  Monica Gotta  Patrick Meraldi 《The Journal of cell biology》2012,199(7):1025-1035
  相似文献   

11.
Tethering Factors Required for Cytokinesis in Arabidopsis     
Martha Thellmann  Katarzyna Rybak  Knut Thiele  Gerhard Wanner  Farhah F. Assaad 《Plant physiology》2010,154(2):720-732
  相似文献   

12.
Immunomodulation by Mesenchymal Stem Cells in Veterinary Species     
Danielle D Carrade  Dori L Borjesson 《Comparative medicine》2013,63(3):207-217
Mesenchymal stem cells (MSC) are adult-derived multipotent stem cells that have been derived from almost every tissue. They are classically defined as spindle-shaped, plastic-adherent cells capable of adipogenic, chondrogenic, and osteogenic differentiation. This capacity for trilineage differentiation has been the foundation for research into the use of MSC to regenerate damaged tissues. Recent studies have shown that MSC interact with cells of the immune system and modulate their function. Although many of the details underlying the mechanisms by which MSC modulate the immune system have been defined for human and rodent (mouse and rat) MSC, much less is known about MSC from other veterinary species. This knowledge gap is particularly important because the clinical use of MSC in veterinary medicine is increasing and far exceeds the use of MSC in human medicine. It is crucial to determine how MSC modulate the immune system for each animal species as well as for MSC derived from any given tissue source. A comparative approach provides a unique translational opportunity to bring novel cell-based therapies to the veterinary market as well as enhance the utility of animal models for human disorders. The current review covers what is currently known about MSC and their immunomodulatory functions in veterinary species, excluding laboratory rodents.Abbreviations: AT, adipose tissue; BM, Bone marrow; CB, umbilical cord blood; CT, umbilical cord tissue; DC, dendritic cell; IDO, indoleamine 2;3-dioxygenase; MSC, mesenchymal stem cells; PGE2, prostaglandin E2; VEGF, vascular endothelial growth factorMesenchymal stem cells (MSC, alternatively known as mesenchymal stromal cells) were first reported in the literature in 1968.39 MSC are thought to be of pericyte origin (cells that line the vasculature)21,22 and typically are isolated from highly vascular tissues. In humans and mice, MSC have been isolated from fat, placental tissues (placenta, Wharton jelly, umbilical cord, umbilical cord blood), hair follicles, tendon, synovial membrane, periodontal ligament, and every major organ (brain, spleen, liver, kidney, lung, bone marrow, muscle, thymus, pancreas, skin).23,121 For most current clinical applications, MSC are isolated from adipose tissue (AT), bone marrow (BM), umbilical cord blood (CB), and umbilical cord tissue (CT; 11,87,99 Clinical trials in human medicine focus on the use of MSC both for their antiinflammatory properties (graft-versus-host disease, irritable bowel syndrome) and their ability to aid in tissue and bone regeneration in combination with growth factors and bone scaffolds (clinicaltrials.gov).131 For tissue regeneration, the abilities of MSC to differentiate and to secrete mediators and interact with cells of the immune system likely contribute to tissue healing (Figure 1). The current review will not address the specific use of MSC for orthopedic applications and tissue regeneration, although the topic is covered widely in current literature for both human and veterinary medicine.57,62,90

Table 1.

Tissues from which MSC have been isolated
Tissue source (reference no.)
SpeciesFatBone marrowCord bloodCord tissueOther
Cat1348356
Chicken63
Cow13812108
Dog973, 5978, 119139Periodontal ligament65
Goat66964
Horse26, 13037, 40, 12367130Periodontal ligament and gingiva88
Nonhuman primate28, 545
Pig1351147014, 20, 91
Rabbit1288032Fetal liver93
Sheep849542, 55
Open in a separate windowOpen in a separate windowFigure 1.The dual roles of MSC: differentiation and modulation of inflammation.Long-term studies in veterinary species have shown no adverse effects with the administration of MSC in a large number of animals.9,10,53 Smaller, controlled studies on veterinary species have shown few adverse effects, such as minor localized inflammation after MSC administration in vivo.7,15,17,45,86,92,98 Private companies, educational institutions, and private veterinary clinics (including Tufts University, Cummins School of Veterinary Medicine, University of California Davis School of Veterinary Medicine, VetStem, Celavet, Alamo Pintado Equine Medical Center, and Rood and Riddle Equine Hospital) offer MSC as a clinical treatment for veterinary species. Clinical uses include tendon and cartilage injuries, tendonitis, and osteoarthritis and, to a lesser extent, bone regeneration, spinal cord injuries, and liver disease in both large and small animals.38,41,113 Even with this broad clinical use, there have been no reports of severe adverse effects secondary to MSC administration in veterinary patients.  相似文献   

13.
The Golgi and the centrosome: building a functional partnership     
Christine Sütterlin  Antonino Colanzi 《The Journal of cell biology》2010,188(5):621-628
The mammalian Golgi apparatus is characterized by a ribbon-like organization adjacent to the centrosome during interphase and extensive fragmentation and dispersal away from the centrosome during mitosis. It is not clear whether this dynamic association between the Golgi and centrosome is of functional significance. We discuss recent findings indicating that the Golgi–centrosome relationship may be important for directional protein transport and centrosome positioning, which are both required for cell polarization. We also summarize our current knowledge of the link between Golgi organization and cell cycle progression.

Introduction

The Golgi apparatus plays a central role in the secretory pathway. Newly synthesized proteins are transported from the ER to the Golgi, where they are posttranslationally modified. They are sorted into carriers for delivery to the plasma membrane or the endosomal–lysosomal system. The basic structural unit of the Golgi apparatus is a stack of flattened cisternae that is morphologically conserved among most species. In mammalian cells, individual Golgi stacks are connected laterally to form a continuous membranous system called the Golgi ribbon, which is located in close physical proximity to the centrosome (Fig. 1, left).Open in a separate windowFigure 1.The spatial relationship between the Golgi and the centrosome during the mammalian cell cycle. Golgi (red, stained with antibodies to GM130) and centrosome (green, stained with antibodies to centrin) staining of nonsynchronized bone cancer cells (U2-OS) shows the physical proximity between these two organelles during interphase (left) and its temporary loss during mitosis (right). Bar, 10 µm.The centrosome functions as the major microtubule-organizing center of the cell and plays an important role in cell polarization and ciliogenesis (Bettencourt-Dias and Glover, 2007). In a newly formed daughter cell, this nonmembrane-bound organelle is composed of a pair of centrioles that is surrounded by a cloud of electron-dense material called the pericentriolar matrix. γ-Tubulin ring complexes (γ-TuRCs) in the pericentriolar matrix allow the centrosome to nucleate the radial array of interphase microtubules whose minus ends are embedded in the centrosome and whose plus ends extend toward the cell periphery. After centrosome duplication in S phase, the two centrosomes move to opposite poles of the cell and become the spindle poles from which spindle microtubules grow. Centrosomes are generally located in the cell center close to the nucleus, although this central position is lost in response to a polarization stimulus, which prompts centrosomes to reorient toward the leading edge of the cell (Pouthas et al., 2008). In most cell types, centrosome reorientation is critical for the ability of cells to polarize and migrate (Yvon et al., 2002). The centrosome is also linked to ciliogenesis because one of its centrioles is converted into the basal body from which a primary cilium extends (D''Angelo and Franco, 2009).The spatial relationship between the Golgi apparatus and the centrosome is altered by changes in Golgi organization that occur during the cell cycle (Fig. 1). These two organelles are only adjacent in interphase when the Golgi apparatus is arranged as a ribbon in the pericentriolar region (Colanzi et al., 2003). In contrast, Golgi membranes are fragmented and dispersed throughout the cytoplasm during mitosis. Intriguingly, the pericentriolar localization of the Golgi is a feature typical of some eukaryotic cells, ranging from mammalian and amphibian cells (Thyberg and Moskalewski, 1999; Reilein et al., 2003) to amoeba (Rehberg et al., 2005). However, other eukaryotes, including plants and flies, have isolated Golgi stacks (Stanley et al., 1997; Nebenführ and Staehelin, 2001) or isolated cisternae in the case of Saccharomyces cerevisiae that are scattered throughout the cytoplasm without an obvious connection with the centrosome (Preuss et al., 1992).In this paper, we review recent findings indicating that the relationship between the Golgi and the centrosome in interphase is important for cell polarization. We also summarize the current understanding of how Golgi–centrosome interactions during mitosis affect cell division.

Are there functional interactions between the Golgi and the centrosome during interphase?

Golgi membranes are actively positioned in the pericentriolar position.

The localization of the mammalian Golgi ribbon next to the centrosome requires the microtubule and actin cytoskeleton (Brownhill et al., 2009). Microtubules have a dual role in organizing the pericentriolar Golgi ribbon. First, the subset of microtubules that is nucleated at the Golgi is necessary for the assembly of Golgi fragments into a connected ribbon in the cell periphery (Miller et al., 2009). Second, centrosomal microtubules provide the tracks along which Golgi membranes are transported to the cell center (Cole et al., 1996). Both steps depend on the minus end–directed motor complex dynein (Burkhardt et al., 1997; Miller et al., 2009). The actin cytoskeleton is also involved in localizing Golgi membranes. Actin fibers, which have been detected at the Golgi complex, are required for the maintenance of the pericentriolar position of this organelle by providing tracks for actin-based motors (Valderrama et al., 1998; Sahlender et al., 2005; Vicente-Manzanares et al., 2007). Actin fibers and microtubules are coordinated by proteins that associate with both cytoskeletal elements such as WHAMM, a Golgi-bound actin-nucleating factor, and MACF1, a microtubule–actin cross-linking protein (Lin et al., 2005; Campellone et al., 2008).Golgi organization and localization in the pericentriolar region also depend on Golgi-associated proteins (Ramirez and Lowe, 2009). Their depletion produces defects in Golgi organization ranging from a disconnected Golgi ribbon in the pericentriolar region (Puthenveedu et al., 2006) to dispersed ministacks in the cytoplasm (Diao et al., 2003; Yadav et al., 2009).

Table I.

Golgi-associated proteins that control the pericentriolar position of the Golgi apparatus
Golgi-associated proteinsReference
Structural Golgi proteins
Cog3Zolov and Lupashin, 2005
GCC185Derby et al., 2007
GCP60 (ACBD3)Sohda et al., 2001
GM130Marra et al., 2007
Golgin-45Short et al., 2001
Golgin-84Diao et al., 2003
Golgin-97Lu et al., 2004
Golgin-160Yadav et al., 2009
Golgin-245Yoshino et al., 2005
GRASP55Feinstein and Linstedt, 2008
GRASP65Puthenveedu et al., 2006
p115Sohda et al., 2005
Membrane traffic
RINT-1Sun et al., 2007
Syntaxin 5Suga et al., 2005
ZW10Sun et al., 2007
Cytoskeleton regulators and motors
ARHGAP10Dubois et al., 2005
CG-NAP/AKAP450Takahashi et al., 1999
CLASP2Efimov et al., 2007
Coronin 7Rybakin et al., 2006
GMAP-210Ríos et al., 2004
FTCDGao and Sztul, 2001
Hook3Walenta et al., 2001
MACF1bLin et al., 2005
Myosin IIVicente-Manzanares et al., 2007
Myosin VISahlender et al., 2005
OptineurinSahlender et al., 2005
p50/dynamitinRoghi and Allan, 1999
WHAMMCampellone et al., 2008
Kinases and enzymes
Cdk5 (kinase)Sun et al., 2008
ORP9 (lipid transfer)Ngo and Ridgway, 2009
PKA (kinase)Bejarano et al., 2006
PKD1 (kinase)Díaz Añel and Malhotra, 2005
Sac1 (PI phosphatase)Liu et al., 2008
Open in a separate windowAlthough the position of the Golgi next to the centrosome is actively maintained, it does not appear to be critical for basic Golgi functions. For example, membrane trafficking and the modification of secretory proteins are unaffected when the Golgi ribbon is severed into individual ministacks (Cole et al., 1996; Diao et al., 2003; Yadav et al., 2009). Furthermore, organisms such as S. cerevisiae secrete proteins with high efficiency, although their Golgi membranes are never pericentriolar (Preuss et al., 1992). Thus, the physiological role of the pericentrosomal positioning of the mammalian Golgi apparatus remains a major unanswered question.

An emerging role for Golgi–centrosome association in polarized secretion.

New studies indicate that the relationship between the Golgi and the centrosome may be important for specialized functions of mammalian cells. A prominent example is cell polarization, which is a prerequisite for cell migration (Li et al., 2005). Cell polarization depends on directional protein transport along Golgi-nucleated microtubules as well as centrosome reorientation toward the leading edge of the cell, which both appear to be affected by interactions between the Golgi and the centrosome.In a recent study, Yadav et al. (2009) investigated the role of the pericentriolar Golgi ribbon in directional transport and cell polarization. Depletion of each of the two structural proteins of the golgin family, GMAP210 and Golgin-160, disrupted the ribbon-like structure of the Golgi and led to isolated ministacks in the cytoplasm. These dispersed stacks were competent of general protein transport to the cell surface. However, there were defects in directional protein transport, as shown by the failure to secrete vesicular stomatitis virus G protein in a directional manner toward the leading edge of a cell and the inability of these cells to migrate in a wound-healing assay. These results indicate that the pericentriolar Golgi ribbon is critical for directional protein transport, although it is not clear whether it is the ribbon-like organization or the position next to the centrosome that is important.Golgi–centrosome interactions may also contribute to cell polarization through regulatory effects on centrosome positioning. Both the centrosome and the Golgi apparatus undergo reorientation toward the leading edge of a stimulated cell. Bisel et al. (2008) found that centrosome reorientation depends on phosphorylation of the Golgi protein GRASP65, which is proposed to promote Golgi stack disassembly (Wang et al., 2003; Yoshimura et al., 2005). In this study, expression of nonphosphorylatable forms of GRASP65 prevented Golgi and centrosome reorientation toward the leading edge and cell migration. Intriguingly, this block was overcome when Golgi membranes were artificially fragmented, indicating that Golgi membranes have to be remodeled to allow the coordinated reorientation of the centrosome and the Golgi. Thus, the ability of the Golgi to reorganize affects the positioning of the centrosome (Bisel et al., 2008).The peripheral Golgi protein, GM130, is an additional critical factor in the regulation of cell polarization (Preisinger et al., 2004; Kodani et al., 2009; Rivero et al., 2009). There are at least four reasons to explain why depletion of GM130 prevents cells from polarizing and migrating in wound-healing assays (Kodani et al., 2009). First, Kodani and Sütterlin (2008) showed that GM130 depletion altered the organization of the centrosome so that it was no longer able to nucleate microtubules or to reorient in response to a polarization stimulus. Second, GM130-dependent centrosome regulation involved the small GTPase Cdc42 (Kodani et al., 2009), a known regulator of cell polarization (Etienne-Manneville, 2006; Kodani et al., 2009). Third, Rivero et al. (2009) identified a novel role for GM130 in microtubule nucleation at the Golgi, which required GM130-dependent recruitment of the microtubule nucleation factor AKAP450 to the Golgi (Rivero et al., 2009). Golgi-nucleated microtubules, which were first identified in in vitro studies (Chabin-Brion et al., 2001), are preferentially oriented toward the leading edge of a motile cell and are necessary for directional protein transport (Fig. 2; Rivero et al., 2009). Fourth, GM130 binds and activates the protein kinase YSK1, which has a known role in cell migration (Preisinger et al., 2004). Thus, GM130 may affect cell polarization and migration through effects on centrosome organization, Cdc42 activation, microtubule nucleation at the Golgi, and YSK1 activation.Open in a separate windowFigure 2.Golgi- and centrosome-nucleated microtubules in cell migration. The centrosome nucleates a radial array of microtubules (red) whose minus ends (−) are anchored at the centrosome and whose plus ends (+) extend into the cell periphery. This population of microtubules depends on γ-TuRC complexes and the large scaffold protein AKAP450 for their nucleation and functions in maintaining the pericentriolar localization of the Golgi ribbon by a dynein-mediated mechanism (closed arrows). In contrast, the Golgi apparatus nucleates microtubules (brown) that extend asymmetrically toward the leading edge of a migrating cell. Microtubule nucleation at the Golgi requires the peripheral Golgi protein GM130, which recruits AKAP450 and γ-TuRC complexes to the Golgi apparatus. Golgi-nucleated microtubules are coated with CLASP proteins and are necessary for the formation of the Golgi ribbon from dispersed stacks. In addition, they are required for cell migration by facilitating polarized protein transport to the leading edge of a cell (open arrows).The formation of a primary cilium is another process that involves interactions between the Golgi and the centrosome. During ciliogenesis, the centrosome moves to the plasma membrane, where one of its centrioles becomes the basal body from which the primary cilium extends. IFT20, a critical component of the intraflagellar transport machinery that is required for formation and extension of the cilium (Follit et al., 2006), localizes to the Golgi by binding to the structural Golgi protein GMAP210. Loss of either IFT20 or GMAP210 impairs ciliogenesis (Follit et al., 2006, 2008), which supports a role for Golgi-localized IFT20 in protein sorting at the Golgi to produce transport carriers involved in the formation of a primary cilium. A similar role in directing specific cargo molecules to the ciliary membrane has been proposed for the small GTPase Rab8, which also localizes to the Golgi and the basal body (Nachury et al., 2007). Collectively, these new findings are intriguing, as they provide support for a functional link between the Golgi and the centrosome.

Are there functional interactions between the Golgi and the centrosome during mitosis?

Regulation of mitotic Golgi reorganization from the centrosome.

The physical proximity of the Golgi apparatus and the centrosome is transiently lost during mitosis when Golgi membranes undergo extensive fragmentation. This dramatic change in Golgi structure is concomitant with a block in secretory trafficking and the reorganization of the microtubule cytoskeleton (Colanzi et al., 2003). Although the Golgi and the centrosome are physically separate at this stage of the cell cycle, there is evidence for functional interactions between these two organelles, which may control progression through mitosis.Many studies have identified possible links between mitotic Golgi fragmentation and the centrosome. For instance, breaking the Golgi ribbon into its constituent stacks during G2 requires the activity of the protein kinase Plk3 (Xie et al., 2004; López-Sánchez et al., 2009), which localizes to the centrosome and spindle poles (Xie et al., 2004; Jiang et al., 2006). The subsequent conversion of Golgi stacks into small, highly dispersed fragments (Jesch et al., 2001; Altan-Bonnet et al., 2006) and vesicular/tubular clusters next to astral spindle microtubules (Shima et al., 1998; Wei and Seemann, 2009) is regulated by two mitotic kinases, Cdk1 and Plk1, which are both associated with the centrosome (Fig. 3; Bailly et al., 1989; Dai and Cogswell, 2003). These findings suggest that components of the centrosome, spindle poles, or the spindle may initiate a signaling pathway that leads to the fragmentation of the Golgi and that may help coordinate Golgi dynamics with cell cycle progression. However, these regulatory factors also exist in the cytosol, and possible roles of cytosolic pools of Cdk1 and Plk3 in mitotic Golgi fragmentation have not been excluded.Open in a separate windowFigure 3.Golgi fragmentation during mitosis. The mammalian Golgi apparatus (green) forms an interconnected ribbon adjacent to the centrosome (red) and the nucleus (blue). It nucleates a population of microtubules that is necessary for polarized protein transport. Plus (+) and minus ends (−) are indicated. The activities of the protein kinases Plk3 and MEK1 and the fission protein BARS are required to convert the ribbon structure into isolated stacks in late G2 and prophase. In metaphase, the isolated stacks are further fragmented by a Plk1- and Cdk1-dependent mechanism, producing vesicular/tubular membranes that are dispersed throughout the cytoplasm. During this process, ribbon determinants, which are proteins required for postmitotic Golgi ribbon formation, remain associated with the mitotic spindle for their partitioning into daughter cells. Centrosome-associated regulators of mitotic Golgi fragmentation are labeled in red. Regulators of Golgi fragmentation that are not associated with the centrosome are labeled in black.Further support for functional interactions between the Golgi and the centrosome during mitosis stems from a novel study on spindle-dependent reassembly of the Golgi ribbon after mitosis (Wei and Seemann, 2009). In a series of elegant experiments, Wei and Seemann (2009) demonstrated that the spindle is required for the postmitotic reformation of the Golgi ribbon. They induced asymmetric cell division so that the entire spindle segregated into only one daughter cell. Although Golgi membranes assembled into stacks in both daughter cells, they only formed a ribbon in the cell that inherited the spindle. Ribbon formation in the spindle-free cell required coinjection of Golgi extracts and tubulin or the addition of spindle-containing fractions. Collectively, these results suggest that Golgi ribbon formation occurs in two steps, with the initial assembly into stacks being mediated by factors that are partitioned by a spindle-independent mechanism. The subsequent formation of the Golgi ribbon from individual stacks, however, has an additional requirement for ribbon determinants, which are likely to be Golgi-associated proteins inherited with the spindle. Possible candidates include regulators of Golgi dynamics and the secretory pathway that have been identified in preparations of the spindle matrix (Ma et al., 2009).

Significance of the loss of Golgi–centrosome proximity during mitosis.

Several studies have identified an unexpected link between mitotic Golgi fragmentation and cell cycle progression (Sütterlin et al., 2002; Hidalgo Carcedo et al., 2004; Preisinger et al., 2005). For example, interfering with mitotic Golgi disassembly by blocking the function of the peripheral Golgi protein GRASP65 or the fission protein BARS resulted in cell cycle arrest in G2 (Sütterlin et al., 2002; Hidalgo Carcedo et al., 2004). Intriguingly, breaking the ribbon into isolated stacks, which occurs in G2, is sufficient to overcome this cell cycle arrest and allows cells to enter mitosis (Colanzi et al., 2007; Feinstein and Linstedt, 2007). It is not known how and why the presence of an intact pericentriolar Golgi ribbon prevents mitotic entry. The existence of a Golgi checkpoint, which monitors the correct inheritance of the Golgi complex, has been proposed because these inhibitory effects are not caused by activation of the DNA damage checkpoint (Sütterlin et al., 2002; Hidalgo Carcedo et al., 2004). It is conceivable that severing the Golgi ribbon in G2 separates ribbon determinants from the rest of the Golgi so that they can cosegregate with the spindle (Fig. 3). Such a mechanism for spindle-dependent Golgi inheritance would ensure that both daughter cells inherit the ability to form a Golgi ribbon and, thus, to transport proteins in a polarized manner. By analogy to the spindle checkpoint, which controls the exit from mitosis by monitoring the correct binding of spindle microtubules to kinetochores, this Golgi checkpoint may assess binding of spindle microtubules to these putative ribbon determinants to regulate entry into mitosis.

Golgi-dependent regulation of the spindle and mitotic progression.

A series of recent studies has identified a requirement for specific Golgi-associated proteins in the formation of a bipolar spindle (Chang et al., 2005), the putative Golgi stacking factor, GRASP65 (Sütterlin et al., 2005), a regulator of the spindle checkpoint, RINT-1 (Lin et al., 2007), and the phosphatidylinositide phosphatase, Sac1 (Burakov et al., 2008). Depletion of any one of these proteins leads to multipolar spindles and mitotic cell death. For example, RNAi-mediated knockdown of Sac1 resulted in disorganization of the Golgi apparatus and mitotic defects characterized by multiple mechanically active spindles (Liu et al., 2008). Similarly, loss of GRASP65 led to the formation of multipolar spindles and mitotic arrest followed by cell death (Sütterlin et al., 2005). The molecular mechanisms by which Golgi-associated proteins regulate spindle formation are not known. Also, it is not known whether Golgi components control spindle formation when Golgi membranes are in the form of a pericentriolar ribbon, isolated stacks, or small fragments.

Table II.

Golgi-associated proteins with a role in regulating centrosome and spindle function
ProteinFunctionDepletion phenotypeReference
Sac1Lipid phosphataseMultiple mechanically active spindlesLiu et al., 2008
GM130GolginAberrant centrosome, multipolar spindlesKodani and Sütterlin, 2008
GRASP65Golgi matrixMultipolar spindles, mitotic cell deathSütterlin et al., 2005
RINT-1Membrane trafficMultipolar spindles, mitotic cell deathSun et al., 2007
Tankyrase-1ADP ribosyl transferaseMultipolar spindles, mitotic cell deathChang et al., 2005
Rab6′GTPaseMetaphase block, SAC activationMiserey-Lenkei et al., 2006
ClathrinVesicle coatDefects in chromosome congression, SAC activationRoyle et al., 2005
Open in a separate windowSAC, spindle assembly checkpoint.In addition to Golgi-dependent effects on spindle formation, other mitotic events are also regulated by disassembly of Golgi stacks during prophase and prometaphase. Indeed, this disassembly step correlates with the release of several peripheral proteins from Golgi membranes to carry out specific functions during mitosis. For instance, clathrin dissociates from the Golgi complex and from endocytic vesicles during mitosis and localizes to the spindle pole where it stabilizes mitotic spindle fibers involved in chromosome segregation (Royle et al., 2005). The small GTPase, Rab6A′, is also released from the Golgi during mitotic Golgi fragmentation (Miserey-Lenkei et al., 2006). If this dynamic behavior of Rab6A′ is inhibited, cells are no longer able to progress through mitosis and are blocked in metaphase through activation of the spindle checkpoint. Another example is the Golgi-associated protein ACBD3, whose release and cytoplasmic dispersal during mitotic Golgi breakdown is necessary for the activation of Numb in the regulation of asymmetric cell division (Zhou et al., 2007). Thus, in addition to facilitating the partitioning of Golgi membranes into the daughter cells, Golgi fragmentation may provide a unique mechanism for the regulation of signaling pathways that involve Golgi-associated components. In the case of ACBD3 and Rab6A′, Golgi fragmentation may relieve inhibitory effects that are either the result of proximity with the centrosome or the organization of the Golgi ribbon.

Conclusions

There is increasing evidence that the relationship between the Golgi apparatus and the centrosome in mammalian cells extends beyond physical proximity and involves functional interactions. Several features of this Golgi–centrosome relationship can be surmised from the recent studies reviewed. This relationship appears to be bidirectional because components of each organelle are able to influence the function of the other organelle. For example, Golgi proteins are necessary for centrosome organization and positioning (Chang et al., 2005; Sütterlin et al., 2005; Kodani and Sütterlin, 2008), whereas centrosome-nucleated microtubules are required for pericentriolar Golgi positioning (Corthésy-Theulaz et al., 1992; Cole et al., 1996). Importantly, these functional interactions affect fundamental cellular processes such as cell polarization and progression through mitosis (Sütterlin et al., 2002; Yadav et al., 2009). Intriguingly, there is evidence for functional interactions when the Golgi and the centrosome are in physical proximity during interphase but also during mitosis when they are physically separate.What is the functional significance of the physical proximity between the Golgi and the centrosome? One possibility is that it may enhance the efficiency of signaling between the Golgi and centrosome and thereby facilitate directional protein transport. The Golgi apparatus is well known for its role in the exocytic pathway, and Golgi membranes, the intermediate compartment, and late endosomes are concentrated in the centrosomal area in mammalian cells (Marie et al., 2009). Thus, the centrosomal area may serve as a traffic hub, allowing integrated regulation of exocytic and endocytic transport routes for polarized delivery of cargo. In support of this idea, species in which Golgi membranes are not adjacent to the centrosome use alternative strategies for transporting proteins in a directional manner. For example, polarized secretion in Drosophila melanogaster is achieved by targeting mRNA to specific transitional ER–Golgi units in which the cargo is synthesized and secreted locally (Herpers and Rabouille, 2004).Why has it taken so long to reveal functional Golgi–centrosome interactions during cell division? The phenomenon of a pericentriolar interphase Golgi ribbon, which is fragmented and dispersed during mitosis, is mainly seen in mammalian cells. Therefore, the significance of this dynamic spatial relationship cannot be studied in a more genetically tractable system such as yeast or Drosophila in which genome-wide screens can be readily performed. Furthermore, there has been a lack of tools to separate the Golgi and centrosome without affecting the functions of these organelles. Some recent studies have used new approaches such as severing the Golgi ribbon by depleting structural golgins, but there are still experimental limitations. For example, an intact Golgi ribbon cannot simply be displaced from the pericentriolar region, which makes it difficult to directly test the significance of Golgi localization versus organization. In addition, Golgi fragmentation, as induced by the depletion of structural Golgi proteins, is a multifactorial process that is marked by both the loss of the Golgi ribbon and dispersal from the pericentriolar position. The limited availability of experimental tools makes it difficult to separate these processes, which has hampered efforts to dissect their individual contributions to the Golgi–centrosome partnership. Also, until a recent study (Kodani et al., 2009), a molecular pathway linking the Golgi and the centrosome during interphase had not been described. For these reasons, it has been difficult to experimentally alter Golgi–centrosome proximity and assay the effects.Although progress has been made, there are many unresolved questions about the Golgi–centrosome relationship during the cell cycle. For example, is there a single bidirectional regulatory pathway between the Golgi and the centrosome, or are there separate signaling pathways in each direction? Are there differences in signaling between these organelles during interphase when the organelles are adjacent and in mitosis when they are physically separate? There are also more specific unanswered questions. For example, how do Golgi proteins control spindle formation? Which factors on the mitotic spindle regulate postmitotic reassembly of the Golgi? How does the organization of the Golgi apparatus control progression through the cell cycle? Is there a Golgi organization checkpoint, and what does it monitor? The answers to these questions will help us better understand the significance of Golgi–centrosome interactions and could lead to the development of novel approaches for the treatment of several important diseases, including cancer.  相似文献   

14.
Poly(ADP-ribose): An organizer of cellular architecture     
Anthony K.L. Leung 《The Journal of cell biology》2014,205(5):613-619
  相似文献   

15.
Arabidopsis LON2 Is Necessary for Peroxisomal Function and Sustained Matrix Protein Import     
Matthew J. Lingard  Bonnie Bartel 《Plant physiology》2009,151(3):1354-1365
Relatively little is known about the small subset of peroxisomal proteins with predicted protease activity. Here, we report that the peroxisomal LON2 (At5g47040) protease facilitates matrix protein import into Arabidopsis (Arabidopsis thaliana) peroxisomes. We identified T-DNA insertion alleles disrupted in five of the nine confirmed or predicted peroxisomal proteases and found only two—lon2 and deg15, a mutant defective in the previously described PTS2-processing protease (DEG15/At1g28320)—with phenotypes suggestive of peroxisome metabolism defects. Both lon2 and deg15 mutants were mildly resistant to the inhibitory effects of indole-3-butyric acid (IBA) on root elongation, but only lon2 mutants were resistant to the stimulatory effects of IBA on lateral root production or displayed Suc dependence during seedling growth. lon2 mutants displayed defects in removing the type 2 peroxisome targeting signal (PTS2) from peroxisomal malate dehydrogenase and reduced accumulation of 3-ketoacyl-CoA thiolase, another PTS2-containing protein; both defects were not apparent upon germination but appeared in 5- to 8-d-old seedlings. In lon2 cotyledon cells, matrix proteins were localized to peroxisomes in 4-d-old seedlings but mislocalized to the cytosol in 8-d-old seedlings. Moreover, a PTS2-GFP reporter sorted to peroxisomes in lon2 root tip cells but was largely cytosolic in more mature root cells. Our results indicate that LON2 is needed for sustained matrix protein import into peroxisomes. The delayed onset of matrix protein sorting defects may account for the relatively weak Suc dependence following germination, moderate IBA-resistant primary root elongation, and severe defects in IBA-induced lateral root formation observed in lon2 mutants.Peroxisomes are single-membrane-bound organelles found in most eukaryotes. Peroxin (PEX) proteins are necessary for various aspects of peroxisome biogenesis, including matrix protein import (for review, see Distel et al., 1996; Schrader and Fahimi, 2008). Most matrix proteins are imported into peroxisomes from the cytosol using one of two targeting signals, a C-terminal type 1 peroxisome-targeting signal (PTS1) or a cleavable N-terminal type 2 peroxisome-targeting signal (PTS2) (Reumann, 2004). PTS1- and PTS2-containing proteins are bound in the cytosol by soluble matrix protein receptors, escorted to the peroxisome membrane docking complex, and translocated into the peroxisome matrix (for review, see Platta and Erdmann, 2007). Once in the peroxisome, many matrix proteins participate in metabolic pathways, such as β-oxidation, hydrogen peroxide decomposition, and photorespiration (for review, see Gabaldon et al., 2006; Poirier et al., 2006).In addition to metabolic enzymes, several proteases are found in the peroxisome matrix. Only one protease, DEG15/Tysnd1, has a well-defined role in peroxisome biology. The rat Tysnd1 protease removes the targeting signal after PTS2-containing proteins enter the peroxisome and also processes certain PTS1-containing β-oxidation enzymes (Kurochkin et al., 2007). Similarly, the Arabidopsis (Arabidopsis thaliana) Tysnd1 homolog DEG15 (At1g28320) is a peroxisomal Ser protease that removes PTS2 targeting signals (Helm et al., 2007; Schuhmann et al., 2008).In contrast with DEG15, little is known about the other eight Arabidopsis proteins that are annotated as proteases in the AraPerox database of putative peroxisomal proteins (Reumann et al., 2004; Carter et al., 2004; Shimaoka et al., 2004), which, in combination with the minor PTS found in both of these predicted proteases (Reumann, 2004), suggests that these enzymes may not be peroxisomal. Along with DEG15, only two of the predicted peroxisomal proteases, an M16 metalloprotease (At2g41790), which we have named PXM16 for peroxisomal M16 protease, and a Lon-related protease (At5g47040/LON2; Ostersetzer et al., 2007), are found in the proteome of peroxisomes purified from Arabidopsis suspension cells (Eubel et al., 2008). DEG15 and LON2 also have been validated as peroxisomally targeted using GFP fusions (Ostersetzer et al., 2007; Schuhmann et al., 2008).

Table I.

Putative Arabidopsis proteases predicted or demonstrated to be peroxisomal
AGI IdentifierAliasProtein ClassT-DNA Insertion AllelesPTSLocalization EvidenceLocalization References
At1g28320DEG15PTS2-processing proteaseSALK_007184 (deg15-1)SKL>aGFPReumann et al., 2004; Helm et al., 2007; Eubel et al., 2008; Schuhmann et al., 2008)
Proteomics
Bioinformatics
At2g41790PXM16Peptidase M16 family proteinSALK_019128 (pxm16-1)PKL>bProteomicsReumann et al., 2004, 2009; Eubel et al., 2008)
SALK_023917 (pxm16-2)Bioinformatics
At5g47040LON2Lon protease homologSALK_128438 (lon2-1)SKL>aGFPReumann et al., 2004, 2009; Ostersetzer et al., 2007; Eubel et al., 2008)
SALK_043857 (lon2-2)Proteomics
Bioinformatics
At2g18080Ser-type peptidaseSALK_020628SSI>cBioinformatics(Reumann et al., 2004)
SALK_102239
At2g35615Aspartyl proteaseSALK_090795ANL>bBioinformatics(Reumann et al., 2004)
SALK_036333
At3g57810Ovarian tumor-like Cys proteaseSKL>aBioinformatics(Reumann et al., 2004)
At4g14570Acylaminoacyl-peptidase proteinCKL>bBioinformatics (peroxisome)(Reumann et al., 2004; Shimaoka et al., 2004)
Proteomics (vacuole)
At4g20310Peptidase M50 family proteinRMx5HLdBioinformatics(Reumann et al., 2004)
At4g36195Ser carboxypeptidase S28 familySSM>bBioinformatics (peroxisome)(Carter et al., 2004; Reumann et al., 2004)





Proteomics (vacuole)

Open in a separate windowaMajor PTS1 (Reumann, 2004).bMinor PTS1 (Reumann, 2004).cValidated PTS1 (Reumann et al., 2007).dMinor PTS2 (Reumann, 2004).PXM16 is the only one of the nine Arabidopsis M16 (pitrilysin family) metalloproteases (García-Lorenzo et al., 2006; Rawlings et al., 2008) containing a predicted PTS. M16 subfamilies B and C contain the plastid and mitochondrial processing peptidases (for review, see Schaller, 2004), whereas PXM16 belongs to M16 subfamily A, which includes insulin-degrading peptidases (Schaller, 2004). A tomato (Solanum lycopersicum) M16 subfamily A protease similar to insulin-degrading enzymes with a putative PTS1 was identified in a screen for proteases that cleave the wound response peptide hormone systemin (Strassner et al., 2002), but the role of Arabidopsis PXM16 is unknown.Arabidopsis LON2 is a typical Lon protease with three conserved domains: an N-terminal domain, a central ATPase domain in the AAA family, and a C-terminal protease domain with a Ser-Lys catalytic dyad (Fig. 1A; Lee and Suzuki, 2008). Lon proteases are found in prokaryotes and in some eukaryotic organelles (Fig. 1C) and participate in protein quality control by cleaving unfolded proteins and can regulate metabolism by controlling levels of enzymes from many pathways, including cell cycle, metabolism, and stress responses (for review, see Tsilibaris et al., 2006). Four Lon homologs are encoded in the Arabidopsis genome; isoforms have been identified in mitochondria, plastids, and peroxisomes (Ostersetzer et al., 2007; Eubel et al., 2008; Rawlings et al., 2008). Mitochondrial Lon protesases are found in a variety of eukaryotes (Fig. 1A) and function both as ATP-dependent proteases and as chaperones promoting protein complex assemblies (Lee and Suzuki, 2008). LON2 is the only Arabidopsis Lon isoform with a canonical C-terminal PTS1 (SKL-COOH; Ostersetzer et al., 2007) or found in the peroxisome proteome (Eubel et al., 2008; Reumann et al., 2009). Functional studies have been conducted with peroxisomal Lon isoforms found in the proteome of peroxisomes purified from rat hepatic cells (pLon; Kikuchi et al., 2004) and the methylotrophic yeast Hansenula polymorpha (Pln; Aksam et al., 2007). Rat pLon interacts with β-oxidation enzymes, and a cell line expressing a dominant negative pLon variant has decreased β-oxidation activity, displays defects in the activation processing of PTS1-containing acyl-CoA oxidase, and missorts catalase to the cytosol (Omi et al., 2008). H. polymorpha Pln is necessary for degradation of a misfolded, peroxisome-targeted version of dihydrofolate reductase and for degradation of in vitro-synthesized alcohol oxidase in peroxisomal matrix extracts, but does not contribute to degradation of peroxisomally targeted GFP (Aksam et al., 2007).Open in a separate windowFigure 1.Diagram of LON2 protein domains, gene models for LON2, PXM16, DEG15, PED1, PEX5, and PEX6, and phylogenetic relationships of LON family members. A, Organization of the 888-amino acid LON2 protein. Locations of the N-terminal domain conserved among Lon proteins, predicted ATP-binding Walker A and B domains (black circles), active site Ser (S) and Lys (K) residues (asterisks), and the C-terminal Ser-Lys-Leu (SKL) peroxisomal targeting signal (PTS1) are shown (Lee and Suzuki, 2008). B, Gene models for LON2, PXM16, DEG15, PED1, PEX5, and PEX6 and locations of T-DNA insertions (triangles) or missense alleles (arrows) used in this study. Exons are depicted by black boxes, introns by black lines, and untranslated regions by gray lines. C, Phylogenetic relationships among LON homologs. Sequences were aligned using MegAlign (DNAStar) and the ClustalW method. The PAUP 4.0b10 program (Swofford, 2001) was used to generate an unrooted phylogram from a trimmed alignment corresponding to Arabidopsis LON2 residues 400 to 888 (from the beginning of the ATPase domain to the end of the protein). The bootstrap method was performed for 500 replicates with distance as the optimality criterion. Bootstrap values are indicated at the nodes. Predicted peroxisomal proteins have C-terminal PTS1 signals in parentheses and are in light-gray ovals. Proteins in the darker gray oval have N-terminal extensions and include mitochondrial and chloroplastic proteins. Sequence identifiers are listed in Supplemental Table S2.In this work, we examined the roles of several putative peroxisomal proteases in Arabidopsis. We found that lon2 mutants displayed peroxisome-deficient phenotypes, including resistance to the protoauxin indole-3-butyric acid (IBA) and age-dependent defects in peroxisomal import of PTS1- and PTS2-targeted matrix proteins. Our results indicate that LON2 contributes to matrix protein import into Arabidopsis peroxisomes.  相似文献   

16.
Tissue-Specific Expression Patterns of Arabidopsis NF-Y Transcription Factors Suggest Potential for Extensive Combinatorial Complexity          下载免费PDF全文
Nicholas Siefers  Kristen K. Dang  Roderick W. Kumimoto  William Edwards Bynum  IV  Gregory Tayrose  Ben F. Holt  III 《Plant physiology》2009,149(2):625-641
  相似文献   

17.
Natural Infection of Burkholderia pseudomallei in an Imported Pigtail Macaque (Macaca nemestrina) and Management of the Exposed Colony     
Crystal H Johnson  Brianna L Skinner  Sharon M Dietz  David Blaney  Robyn M Engel  George W Lathrop  Alex R Hoffmaster  Jay E Gee  Mindy G Elrod  Nathaniel Powell  Henry Walke 《Comparative medicine》2013,63(6):528-535
Identification of the select agent Burkholderia pseudomallei in macaques imported into the United States is rare. A purpose-bred, 4.5-y-old pigtail macaque (Macaca nemestrina) imported from Southeast Asia was received from a commercial vendor at our facility in March 2012. After the initial acclimation period of 5 to 7 d, physical examination of the macaque revealed a subcutaneous abscess that surrounded the right stifle joint. The wound was treated and resolved over 3 mo. In August 2012, 2 mo after the stifle joint wound resolved, the macaque exhibited neurologic clinical signs. Postmortem microbiologic analysis revealed that the macaque was infected with B. pseudomallei. This case report describes the clinical evaluation of a B. pseudomallei-infected macaque, management and care of the potentially exposed colony of animals, and protocols established for the animal care staff that worked with the infected macaque and potentially exposed colony. This article also provides relevant information on addressing matters related to regulatory issues and risk management of potentially exposed animals and animal care staff.Abbreviations: CDC, Centers for Disease Control and Prevention; IHA, indirect hemagglutination assay; PEP, postexposure prophylacticBurkholderia pseudomallei, formerly known as Pseudomonas pseudomallei, is a gram-negative, aerobic, bipolar, motile, rod-shaped bacterium. B. pseudomallei infections (melioidosis) can be severe and even fatal in both humans and animals. This environmental saprophyte is endemic to Southeast Asia and northern Australia, but it has also been found in other tropical and subtropical areas of the world.7,22,32,42 The bacterium is usually found in soil and water in endemic areas and is transmitted to humans and animals primarily through percutaneous inoculation, ingestion, or inhalation of a contaminated source.8, 22,28,32,42 Human-to-human, animal-to-animal, and animal-to-human spread are rare.8,32 In December 2012, the National Select Agent Registry designated B. pseudomallei as a Tier 1 overlap select agent.39 Organisms classified as Tier 1 agents present the highest risk of deliberate misuse, with the most significant potential for mass casualties or devastating effects to the economy, critical infrastructure, or public confidence. Select agents with this status have the potential to pose a severe threat to human and animal health or safety or the ability to be used as a biologic weapon.39Melioidosis in humans can be challenging to diagnose and treat because the organism can remain latent for years and is resistant to many antibiotics.12,37,41 B. pseudomallei can survive in phagocytic cells, a phenomenon that may be associated with latent infections.19,38 The incubation period in naturally infected animals ranges from 1 d to many years, but symptoms typically appear 2 to 4 wk after exposure.13,17,35,38 Disease generally presents in 1 of 2 forms: localized infection or septicemia.22 Multiple methods are used to diagnose melioidosis, including immunofluorescence, serology, and PCR analysis, but isolation of the bacteria from blood, urine, sputum, throat swabs, abscesses, skin, or tissue lesions remains the ‘gold standard.’9,22,40,42 The prognosis varies based on presentation, time to diagnosis, initiation of appropriate antimicrobial treatment, and underlying comorbidities.7,28,42 Currently, there is no licensed vaccine to prevent melioidosis.There are several published reports of naturally occurring melioidosis in a variety of nonhuman primates (NHP; 2,10,13,17,25,30,31,35 The first reported case of melioidosis in monkeys was recorded in 1932, and the first published case in a macaque species was in 1966.30 In the United States, there have only been 7 documented cases of NHP with B. pseudomallei infection.2,13,17 All of these cases occurred prior to the classification of B. pseudomallei as a select agent. Clinical signs in NHP range from subclinical or subacute illness to acute septicemia, localized infection, and chronic infection. NHP with melioidosis can be asymptomatic or exhibit clinical signs such as anorexia, wasting, purulent drainage, subcutaneous abscesses, and other soft tissue lesions. Lymphadenitis, lameness, osteomyelitis, paralysis and other CNS signs have also been reported.2,7,10,22,28,32 In comparison, human''s clinical signs range from abscesses, skin ulceration, fever, headache, joint pain, and muscle tenderness to abdominal pain, anorexia, respiratory distress, seizures, and septicemia.7,9,21,22

Table 1.

Summary of reported cases of naturally occurring Burkholderia pseudomalleiinfections in nonhuman primates
CountryaImported fromDate reportedSpeciesReference
AustraliaBorneo1963Pongo sp.36
BruneiUnknown1982Orangutan (Pongo pygmaeus)33
France1976Hamlyn monkey (Cercopithecus hamlyni) Patas monkey (Erythrocebus patas)11
Great BritainPhilippines and Indonesia1992Cynomolgus monkey (Macaca fascicularis)10
38
MalaysiaUnknown1966Macaca spp.30
Unknown1968Spider monkey (Brachytelis arachnoides) Lar gibbon (Hylobates lar)20
Unknown1969Pig-tailed macaque (Macaca nemestrina)35
Unknown1984Banded leaf monkey (Presbytis melalophos)25
SingaporeUnknown1995Gorillas, gibbon, mandrill, chimpanzee43
ThailandUnknown2012Monkey19
United StatesThailand1970Stump-tailed macaque (Macaca arctoides)17
IndiaPig-tailed macaque (Macaca nemestrina)
AfricaRhesus macaque (Macaca mulatta) Chimpanzee (Pan troglodytes)
Unknown1971Chimpanzee (Pan troglodytes)3
Malaysia1981Pig-tailed macaque (Macaca nemestrina)2
Wild-caught, unknown1986Rhesus macaque (Macaca mulatta)13
Indonesia2013Pig-tailed macaque (Macaca nemestrina)Current article
Open in a separate windowaCountry reflects the location where the animal was housed at the time of diagosis.Here we describe a case of melioidosis diagnosed in a pigtail macaque (Macaca nemestrina) imported into the United States from Indonesia and the implications of the detection of a select agent identified in a laboratory research colony. We also discuss the management and care of the exposed colony, zoonotic concerns regarding the animal care staff that worked with the shipment of macaques, effects on research studies, and the procedures involved in reporting a select agent incident.  相似文献   

18.
Monkey B Virus (Cercopithecine herpesvirus 1)     
David Elmore  Richard Eberle 《Comparative medicine》2008,58(1):11-21
  相似文献   

19.
Clearance of apoptotic cells: implications in health and disease     
Michael R. Elliott  Kodi S. Ravichandran 《The Journal of cell biology》2010,189(7):1059-1070
  相似文献   

20.
Epistatic Interactions between Opaque2 Transcriptional Activator and Its Target Gene CyPPDK1 Control Kernel Trait Variation in Maize   总被引:1,自引:0,他引:1       下载免费PDF全文
Domenica Manicacci  Letizia Camus-Kulandaivelu  Marie Fourmann  Chantal Arar  Stéphanie Barrault  Agnès Rousselet  No?l Feminias  Luciano Consoli  Lisa Francès  Valérie Méchin  Alain Murigneux  Jean-Louis Prioul  Alain Charcosset  Catherine Damerval 《Plant physiology》2009,150(1):506-520
  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号