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 46 毫秒
1.
In Xenopus egg extracts, spindles assembled around sperm nuclei contain a centrosome at each pole, while those assembled around chromatin beads do not. Poles can also form in the absence of chromatin, after addition of a microtubule stabilizing agent to extracts. Using this system, we have asked (a) how are spindle poles formed, and (b) how does the nucleation and organization of microtubules by centrosomes influence spindle assembly? We have found that poles are morphologically similar regardless of their origin. In all cases, microtubule organization into poles requires minus end–directed translocation of microtubules by cytoplasmic dynein, which tethers centrosomes to spindle poles. However, in the absence of pole formation, microtubules are still sorted into an antiparallel array around mitotic chromatin. Therefore, other activities in addition to dynein must contribute to the polarized orientation of microtubules in spindles. When centrosomes are present, they provide dominant sites for pole formation. Thus, in Xenopus egg extracts, centrosomes are not necessarily required for spindle assembly but can regulate the organization of microtubules into a bipolar array.During cell division, the correct organization of microtubules in bipolar spindles is necessary to distribute chromosomes to the daughter cells. The slow growing or minus ends of the microtubules are focused at each pole, while the plus ends interact with the chromosomes in the center of the spindle (Telzer and Haimo, 1981; McIntosh and Euteneuer, 1984). Current concepts of spindle assembly are based primarily on mitotic spindles of animal cells, which contain centrosomes. Centrosomes are thought to be instrumental for organization of the spindle poles and for determining both microtubule polarity and the spindle axis. In the prevailing model, termed “Search and Capture,” dynamic microtubules growing from two focal points, the centrosomes, are captured and stabilized by chromosomes, generating a bipolar array (Kirschner and Mitchison, 1986). However, while centrosomes are required for spindle assembly in some systems (Sluder and Rieder, 1985; Rieder and Alexander, 1990; Zhang and Nicklas, 1995a ,b), in other systems they appear to be dispensable (Steffen et al., 1986; Heald et al., 1996). Furthermore, centrosomes are not present in higher plant cells and in female meiosis of most animal species (Bajer and Mole, 1982; Gard, 1992; Theurkauf and Hawley, 1992; Albertson and Thomson, 1993; Lambert and Lloyd, 1994). In the absence of centrosomes, bipolar spindle assembly seems to occur through the self-organization of microtubules around mitotic chromatin (McKim and Hawley, 1995; Heald et al., 1996; Waters and Salmon, 1997). The observation of apparently different spindle assembly pathways raises several questions: Do different types of spindles share common mechanisms of organization? How do centrosomes influence spindle assembly? In the absence of centrosomes, what aspects of microtubule self-organization promote spindle bipolarity?To begin to address these questions, we have used Xenopus egg extracts, which can be used to reconstitute different types of spindle assembly. Spindle assembly around Xenopus sperm nuclei is directed by centrosomes (Sawin and Mitchison, 1991). Like other meiotic systems (Bastmeyer et al., 1986; Steffen et al., 1986), Xenopus extracts also support spindle assembly around chromatin in the absence of centrosomes through the movement and sorting of randomly nucleated microtubules into a bipolar structure (Heald et al., 1996). In this process, the microtubule-based motor cytoplasmic dynein forms spindle poles by cross-linking and sliding microtubule minus ends together. Increasing evidence suggests that the function of dynein in spindle assembly depends on its interaction with other proteins, including dynactin, a dynein-binding complex, and NuMA1 (nuclear protein that associates with the mitotic apparatus) (Merdes et al., 1996; Echeverri et al., 1996; Gaglio et al., 1996). In this paper, we demonstrate that both in the presence and absence of centrosomes, spindle pole assembly occurs by a common dynein-dependent mechanism. We show that when centrosomes are present, they are tethered to spindle poles by dynein. In the absence of dynein function, microtubules are still sorted into an antiparallel array, indicating that other aspects of microtubule self-assembly independent of pole formation promote spindle bipolarity around mitotic chromatin. Since centrosomes are dispensable for pole formation in this system, what is their function? We show here that if only one centrosome is present, it acts as a dominant site for microtubule nucleation and focal organization, resulting in a monopolar spindle. Therefore, although centrosomes are not required in this system, they can influence spindle pole formation and bipolarity.  相似文献   

2.
CLIPs (cytoplasmic linker proteins) are a class of proteins believed to mediate the initial, static interaction of organelles with microtubules. CLIP-170, the CLIP best characterized to date, is required for in vitro binding of endocytic transport vesicles to microtubules. We report here that CLIP-170 transiently associates with prometaphase chromosome kinetochores and codistributes with dynein and dynactin at kinetochores, but not polar regions, during mitosis. Like dynein and dynactin, a fraction of the total CLIP-170 pool can be detected on kinetochores of unattached chromosomes but not on those that have become aligned at the metaphase plate. The COOH-terminal domain of CLIP-170, when transiently overexpressed, localizes to kinetochores and causes endogenous full-length CLIP-170 to be lost from the kinetochores, resulting in a delay in prometaphase. Overexpression of the dynactin subunit, dynamitin, strongly reduces the amount of CLIP-170 at kinetochores suggesting that CLIP-170 targeting may involve the dynein/dynactin complex. Thus, CLIP-170 may be a linker for cargo in mitosis as well as interphase. However, dynein and dynactin staining at kinetochores are unaffected by this treatment and further overexpression studies indicate that neither CLIP-170 nor dynein and dynactin are required for the formation of kinetochore fibers. Nevertheless, these results strongly suggest that CLIP-170 contributes in some way to kinetochore function in vivo.Microtubules (MTs)1 in vertebrate somatic cells are involved in intracellular transport and distribution of membranous organelles. Fundamental to this role are their tightly controlled, polarized organization, and unusual dynamic properties (Hirokawa, 1994) and their interaction with a complex set of MT-based motor proteins (Hirokawa, 1996; Sheetz, 1996; Goodson et al., 1997). During mitosis, they contribute to the motility of centrosomes, the construction of spindle poles (Karsenti et al., 1996; Merdes and Cleveland, 1997), and the dynamic movements of kinetochores (Rieder and Salmon, 1994) and chromosome arms (Barton and Goldstein, 1996; Vernos and Karsenti, 1996). The motor protein cytoplasmic dynein, drives the transport toward MT minus-ends of a variety of subcellular organelles (Schnapp and Reese, 1989; Schroer et al., 1989; Holzbaur and Vallee, 1994). Dynactin is a molecular complex originally identified as being essential for dynein-mediated movement of salt-washed vesicles in vitro (reviewed in Schroer, 1996; Schroer and Sheetz, 1991). Genetic studies in fungi, yeast, and flies have shown that the two complexes function together to drive nuclear migration, spindle and nuclear positioning and to permit proper neuronal development (Eshel et al., 1993; Clark and Meyer, 1994; Muhua et al., 1994; Plamann et al., 1994; McGrail et al., 1995; Karsenti et al., 1996). Biochemical studies suggest a direct interaction between certain subunits of dynein and dynactin (Karki and Holzbaur, 1995; Vaughan and Vallee, 1995). In vivo, the two molecules may bind one another transiently, since they have not been isolated as a stable complex.There is good evidence indicating that the dynein/dynactin complex, together with other motors (Eg5, and a minus-end oriented kinesin-related protein) and a structural protein (NuMa), drive the focusing of free microtubule ends into mitotic spindle poles (Merdes and Cleveland, 1997; Waters and Salmon, 1997). A trimolecular complex composed of NuMa and dynein/dynactin may be crucial in this process in both acentriolar (Merdes et al., 1996), and centriolar spindles (Gaglio et al., 1997). A number of findings also indicate that the combined actions of dynein and dynactin at the kinetochore contribute to chromosome alignment in vertebrate somatic cells. First, the initial interaction between polar spindle MTs and kinetochores seems to involve a tangential capture event (Merdes and De Mey, 1990; Rieder and Alexander, 1990) which is followed by a poleward gliding along the surface lattice of the MT (Hayden et al., 1990). Both in vivo and in vitro (Hyman and Mitchison, 1991) this gliding movement appears similar to the dynein-mediated retrograde transport of vesicular organelles along MTs. Consistent with this is the finding that both dynein (Pfarr et al., 1990; Steuer et al., 1990) and its activator, dynactin (Echeverri et al., 1996), are present at prometaphase kinetochores. Overexpression of dynamitin, a 50-kD subunit of the dynactin complex, results in the partial disruption of the dynactin complex and in the loss, from kinetochores, of dynein, as well as dynactin. Therefore, it has been proposed that dynactin mediates the association of dynein with kinetochores. Abnormal spindles with poorly focused poles are observed and the cells become arrested in pseudoprometaphase (Echeverri et al., 1996). Despite these findings, rigorous proof for a role of the dynein motor complex in kinetochore motility is still lacking, and its role may differ between lower and higher eucaryotes, and between mitosis and meiosis.CLIP-170 (Rickard and Kreis, 1996) is needed for in vitro binding of endocytic transport vesicles to MTs (Pierre et al., 1992). It is a nonmotor MT-binding protein that accumulates preferentially in the vicinity of MT plus ends and on early endosomes and endocytic transport vesicles in nondividing cells (Rickard and Kreis, 1990; Pierre et al., 1992). Like many MT-binding proteins, CLIP-170 is a homodimer whose NH2-terminal head domains and COOH-terminal tail domains flank a central α-helical coiled-coil domain. The binding of CLIP-170 to MTs involves a 57–amino acid sequence present twice in the head domain (Pierre et al., 1992) and is regulated by phosphorylation (Rickard and Kreis, 1991). The COOH-terminal domain has been proposed to participate in targeting to endocytic membranes (Pierre et al., 1994). The fact that the latter move predominantly toward microtubule minus ends in a process most likely mediated by cytoplasmic dynein and dynactin (Aniento and Gruenberg, 1995), suggests that CLIP-170 may act in concert with this motor complex, and may be subject to regulated interactions with one or more dynactin or dynein subunits at the vesicle membrane.Here we report that during mitosis, CLIP-170 codistributes with dynein and dynactin at kinetochores, but not spindle poles. Evidence is presented that the COOH-terminal domain of CLIP-170 is responsible for its kinetochore targeting, and that this may be mediated by the complex of dynein and dynactin. The effects on mitotic progression of overexpression of wild type and several deletion mutants of CLIP-170 provide evidence for the involvement of CLIP-170 in kinetochore function early in mitosis. We also present in vivo evidence that neither CLIP-170 nor the complex of dynein and dynactin are required for formation of kinetochore fibers.  相似文献   

3.
The ordered arrangement of cortical microtubules in growing plant cells is essential for anisotropic cell expansion and, hence, for plant morphogenesis. These arrays are dismantled when the microtubule cytoskeleton is rearranged during mitosis and reassembled following completion of cytokinesis. The reassembly of the cortical array has often been considered as initiating from a state of randomness, from which order arises at least partly through self-organizing mechanisms. However, some studies have shown evidence for ordering at early stages of array assembly. To investigate how cortical arrays are initiated in higher plant cells, we performed live-cell imaging studies of cortical array assembly in tobacco (Nicotiana tabacum) Bright Yellow-2 cells after cytokinesis and drug-induced disassembly. We found that cortical arrays in both cases did not initiate randomly but with a significant overrepresentation of microtubules at diagonal angles with respect to the cell axis, which coincides with the predominant orientation of the microtubules before their disappearance from the cell cortex in preprophase. In Arabidopsis (Arabidopsis thaliana) root cells, recovery from drug-induced disassembly was also nonrandom and correlated with the organization of the previous array, although no diagonal bias was observed in these cells. Surprisingly, during initiation, only about one-half of the new microtubules were nucleated from locations marked by green fluorescent protein-γ-tubulin complex protein2-tagged γ-nucleation complexes (γ-tubulin ring complex), therefore indicating that a large proportion of early polymers was initiated by a noncanonical mechanism not involving γ-tubulin ring complex. Simulation studies indicate that the high rate of noncanonical initiation of new microtubules has the potential to accelerate the rate of array repopulation.Higher plant cells feature ordered arrays of microtubules at the cell cortex (Ledbetter and Porter, 1963) that are essential for cell and tissue morphogenesis, as revealed by disruption of cortical arrays by drugs that cause microtubule depolymerization (Green, 1962) or stabilization (Weerdenburg and Seagull, 1988) and by loss-of-function mutations in a wide variety of microtubule-associated proteins (Baskin, 2001; Whittington et al., 2001; Buschmann and Lloyd, 2008; Lucas et al., 2011). The structure of these arrays is thought to control the pattern of cell growth primarily by its role in the deposition of cellulose microfibrils, the load-bearing component of the cell wall (Somerville, 2006). Functional relations between cortical microtubules and cellulose microfibrils have been proposed since the early sixties, even before cortical microtubules had been visualized (Green, 1962). Recent live-cell imaging studies have confirmed that cortical microtubules indeed guide the movement of cellulose synthase complexes that produce cellulose microfibrils (Paredez et al., 2006) and have shown further that microtubules position the insertion of most cellulose synthase complexes into the plasma membrane (Gutierrez et al., 2009). These activities of ordered cortical microtubules are proposed to facilitate the organization of cell wall structure, creating material properties that underlie cell growth anisotropy.While organization of the interphase cortical array appears to be essential for cell morphogenesis, this organization is disrupted during the cell cycle as microtubules are rearranged to create the preprophase band, spindle, and phragmoplast during mitosis and cytokinesis (for review, see Wasteneys, 2002). Upon completion of cytokinesis, an organized interphase cortical array is regenerated, but the pathway for this reassembly is not well understood.The plant interphase microtubule array is organized and maintained without centrosomes as organizing centers (for review, see Wasteneys, 2002; Bartolini and Gundersen, 2006; Ehrhardt and Shaw, 2006), and microtubule self-organization is proposed to play an important role in cortical microtubule array ordering (Dixit and Cyr, 2004). In electron micrographs, microtubules have been observed to be closely associated with the plasma membrane (Hardham and Gunning, 1978), and live-cell imaging provides evidence for attachment of microtubules to the cell cortex (Shaw et al., 2003; Vos et al., 2004). The close association to the plasma membrane restricts the cortical microtubules to a quasi two-dimensional plane where they interact through polymerization-driven collisions (Shaw et al., 2003; Dixit and Cyr, 2004). Microtubule encounters at shallow angles (<40°) have a high probability of leading to bundling, while microtubule encounters at steeper angles most likely result in induced catastrophes or microtubule crossovers (Dixit and Cyr, 2004). Several computational modeling studies have since shown that these types of interactions between surface-bound dynamical microtubules can indeed explain spontaneous coalignment of microtubules (Allard et al., 2010; Eren et al., 2010; Hawkins et al., 2010; Tindemans et al., 2010).The question of how the orientation of the cortical array is established with respect to the cell axis is less well understood. One possibility is that microtubules are selectively destabilized with respect to cellular coordinates (Ehrhardt and Shaw, 2006). Indeed, recent results from biological observations and modeling suggest that catastrophic collisions induced at the edges between cell faces or heighted catastrophe rates in cell caps could be sufficient to selectively favor microtubules in certain orientation and hence determine the final orientation of the array (Allard et al., 2010; Eren et al., 2010; Ambrose et al., 2011; Dhonukshe et al., 2012).To date, all models of cortical array assembly assume random initial conditions. However, experimental work by Wasteneys and Williamson (1989a, 1989b) in Nitella tasmanica showed that, during array reassembly after drug-induced disruption, microtubules were initially transverse. This was followed by a less ordered phase and later by the acquisition of the final transverse order. A nonrandom initial ordering was also observed in tobacco (Nicotiana tabacum) Bright Yellow-2 (BY-2) cells by Kumagai et al. (2001), who concluded that the process of transverse array establishment starts with longitudinal order but did not provide quantitative data for the process of array assembly. The initial conditions for the cortical microtubule array formation are important to consider, as they may strongly influence the speed at which order is established and could even prevent it from being established over a biologically relevant time scale.In this study, we used live-cell imaging to follow and record the whole transition from the cortical microtubule-free state to the final transverse array and used digital tracking algorithms to quantify the microtubule order. Nucleation stands out as a central process to characterize during array initiation. Lacking a central body to organize microtubule nucleations, the higher plant cell has dispersed nucleation complexes (Wasteneys and Williamson, 1989a, 1989b; Chan et al., 2003; Shaw et al., 2003; Murata et al., 2005; Pastuglia et al., 2006; Nakamura et al., 2010). Therefore, we performed high time resolution observations to quantify nucleation complex recruitment, nucleation rates, and microtubule nucleation angles. We found evidence for a highly nonrandom initial ordering state that features diagonal microtubule orientation and an atypical microtubule initiation mechanism. Simulation analysis indicates that these atypical nucleations have the potential to accelerate the recovery of cortical array density.  相似文献   

4.
EMBO J 32: 2905–2919 10.1038/emboj.2013.199; published online September032013Some B cells of the adaptive immune system secrete polyreactive immunoglobulin G (IgG) in the absence of immunization or infection. Owing to its limited affinity and specificity, this natural IgG is thought to play a modest protective role. In this issue, a report reveals that natural IgG binds to microbes following their opsonization by ficolin and mannan-binding lectin (MBL), two carbohydrate receptors of the innate immune system. The interaction of natural IgG with ficolins and MBL protects against pathogenic bacteria via a complement-independent mechanism that involves IgG receptor FcγRI expressing macrophages. Thus, natural IgG enhances immunity by adopting a defensive strategy that crossovers the conventional boundaries between innate and adaptive microbial recognition systems.The adaptive immune system generates protective somatically recombined antibodies through a T cell-dependent (TD) pathway that involves follicular B cells. After recognizing antigen through the B-cell receptor (BCR), follicular B cells establish a cognate interaction with CD4+ T follicular helper (TFH) cells and thereafter either rapidly differentiate into short-lived IgM-secreting plasmablasts or enter the germinal centre (GC) of lymphoid follicles to complete class switch recombination (CSR) and somatic hypermutation (SHM) (Victora and Nussenzweig, 2012). CSR from IgM to IgG, IgA and IgE generates antibodies with novel effector functions, whereas SHM provides the structural correlate for the induction of affinity maturation (Victora and Nussenzweig, 2012). Eventually, this canonical TD pathway generates long-lived bone marrow plasma cells and circulating memory B cells that produce protective class-switched antibodies capable to recognize specific antigens with high affinity (Victora and Nussenzweig, 2012).In addition to post-immune monoreactive antibodies, B cells produce pre-immune polyreactive antibodies in the absence of conventional antigenic stimulation (Ehrenstein and Notley, 2010). These natural antibodies form a vast and stable repertoire that recognizes both non-protein and protein antigens with low affinity (Ehrenstein and Notley, 2010). Natural antibodies usually emerge from a T cell-independent (TI) pathway that involves innate-like B-1 and marginal zone (MZ) B cells. These are extrafollicular B-cell subsets that rapidly differentiate into short-lived antibody-secreting plasmablasts after detecting highly conserved microbial and autologus antigens through polyreactive BCRs and nonspecific germline-encoded pattern recognition receptors (Pone et al, 2012; Cerutti et al, 2013).The most studied natural antibody is IgM, a pentameric complement-activating molecule with high avidity but low affinity for antigen (Ehrenstein and Notley, 2010). In addition to promoting the initial clearance of intruding microbes, natural IgM regulates tissue homeostasis, immunological tolerance and tumour surveillance (Ochsenbein et al, 1999; Zhou et al, 2007; Ehrenstein and Notley, 2010). Besides secreting IgM, B-1 and MZ B cells produce IgG and IgA after receiving CSR-inducing signals from dendritic cells (DCs), macrophages and neutrophils of the innate immune system (Cohen and Norins, 1966; Cerutti et al, 2013). In humans, certain natural IgG and IgA are moderately mutated and show some specificity, which may reflect the ability of human MZ B cells to undergo SHM (Cerutti et al, 2013). Yet, natural IgG and IgA are generally perceived as functionally quiescent.In this issue, Panda et al show that natural IgG bound to a broad spectrum of bacteria with high affinity by cooperating with ficolin and MBL (Panda et al, 2013), two ancestral soluble lectins of the innate immune system (Holmskov et al, 2003). This binding involved some degree of specificity, because it required the presence of ficolin or MBL on the microbial surface as well as lower pH and decreased calcium concentration in the extracellular environment as a result of infection or inflammation (see Figure 1).Open in a separate windowFigure 1Ficolins and MBL are produced by hepatocytes and various cells of the innate immune system and opsonize bacteria after recognizing conserved carbohydrates. Low pH and calcium concentrations present under infection-inflammation conditions promote the interaction of ficolin or MBL with natural IgG on the surface of bacteria. The resulting immunocomplex is efficiently phagocytosed by macrophages through FcγR1 independently of the complement protein C3, leading to the clearance of bacteria.Ficolins and MBL are soluble pattern recognition receptors that opsonize microbes after binding to glycoconjugates through distinct carbohydrate recognition domain (CRD) structures (Holmskov et al, 2003). While ficolins use a fibrinogen domain, MBL and other members of the collectin family use a C-type lectin domain attached to a collagen-like region (Holmskov et al, 2003). Similar to pentraxins, ficolins and MBL are released by innate effector cells and hepatocytes, and thus may have served as ancestral antibody-like molecules prior to the inception of the adaptive immune system (Holmskov et al, 2003; Bottazzi et al, 2010). Of note, MBL and the MBL-like complement protein C1q are recruited by natural IgM to mediate complement-dependent clearance of autologous apoptotic cells and microbes (Holmskov et al, 2003; Ehrenstein and Notley, 2010). Panda et al found that a similar lectin-dependent co-optation strategy enhances the protective properties of natural IgG (Panda et al, 2013).By using bacteria and the bacterial glycan N-acetylglicosamine, Panda et al show that natural IgG isolated from human serum or T cell-deficient mice interacted with the fibrinogen domain of microbe-associated ficolins (Panda et al, 2013). The resulting immunocomplex was phagocytosed by macrophages via the IgG receptor FcγRI in a complement-independent manner (Panda et al, 2013). The additional involvement of MBL was demonstrated by experiments showing that natural IgG retained some bacteria-binding activity in the absence of ficolins (Panda et al, 2013).Surface plasmon resonance provided some clues regarding the molecular requirements of the ficolin–IgG interaction (Panda et al, 2013), but the conformational changes required by ficolin to interact with natural IgG remain to be addressed. In particular, it is unclear what segment of the effector Fc domain of natural IgG binds to ficolins and whether Fc-associated glycans are involved in this binding. Specific glycans have been recently shown to mitigate the inflammatory properties of IgG emerging from TI responses (Hess et al, 2013) and this process could implicate ficolins and MBL. Moreover, it would be important to elucidate whether and how the antigen-binding Fab portion of natural IgG regulates its interaction with ficolins and MBL.The in vivo protective role of natural IgG was elegantly demonstrated by showing that reconstitution of IgG-deficient mice lacking the CSR-enzyme activation-induced cytidine deaminase with natural IgG from T cell-insufficient animals enhanced resistance to pathogenic Pseudomonas aeruginosa (Panda et al, 2013). This protective effect was associated with reduced production of proinflammatory cytokines, occurred independently of the complement protein C3 and was impaired by peptides capable to inhibit the binding of natural IgG to ficolin (Panda et al, 2013). Additional in vivo studies will be needed to determine whether natural IgG exerts protective activity in mice lacking ficolin, MBL or FcγRI, and to ascertain whether these molecules also enhance the protective properties of canonical or natural IgG and IgA released by bone marrow plasma cells and mucosal plasma cells, respectively.In conclusion, the findings by Panda et al show that natural IgG adopts ‘crossover'' defensive strategies that blur the conventional boundaries between the innate and adaptive immune systems. The sophisticated integration of somatically recombined and germline-encoded antigen recognition systems described in this new study shall stimulate immunologists to further explore the often underestimated protective virtues of our vast natural antibody repertoire. This effort may lead to the development of novel therapies against infections.  相似文献   

5.
EMBO J 32 15, 2099–2112 doi:10.1038/emboj.2013.125; published online May312013Mutations in Parkin represent ∼50% of disease-causing defects in autosomal recessive-juvenile onset Parkinson''s disease (AR-JP). Recently, there have been four structural reports of autoinhibited forms of this RING-IBR-RING (RBR) ubiquitin ligase (E3) by the Gehring, Komander, Johnston and Shaw groups. The important advances from these studies set the stage for the next steps in understanding the molecular basis for Parkinson''s disease (PD).Regulated protein degradation requires that E3s and their access to substrates be exquisitely controlled. RBR family E3s provide striking examples of this regulation. The complex and compact structures of Parkin (Riley et al, 2013; Spratt et al, 2013; Trempe et al, 2013; Wauer and Komander, 2013) as well as another RBR E3, human homologue of Ariadne (HHARI) (Duda et al, 2013), demonstrate extraordinarily intricate inter-domain arrangements. These autoinhibited structures ensure that their functions are restricted until activated.Until recently, RBR E3s were believed to be a subclass of RING E3s, which allosterically activate E2 conjugated with ubiquitin (E2∼Ub). However, Wenzel et al (2011) determined that they are actually hybrid E3s, containing an E2 binding site in RING1 and a catalytic cysteine residue in the domain designated as RING2. The catalytic cysteine is an acceptor for an ubiquitin from RING1-bound E2∼Ub forming an intermediate (E3∼Ub) that leads to substrate or autoubiquitination. In this way, RBRs resemble HECT E3s, which also form catalytic intermediates in ubiquitination. There are 13 human RBR family E3s. Besides Parkin, two notable RBRs are HOIL-1 and HOIP, which form part of a complex integral to NF-κB activation (Wenzel and Klevit, 2012).In addition to causal roles in AR-JP, single allele mutations of Parkin are found in some sporadic cases of PD (references in Wauer and Komander, 2013). Mutations in the Parkin-associated kinase PINK1, which is upstream of Parkin, also account for a significant number of AR-JP cases (Hardy et al, 2009; Narendra et al, 2012; Lazarou et al, 2013). A number of diverse Parkin substrates have been postulated to be associated with PD. There is substantial evidence that one role for Parkin is at mitochondria. Once activated and recruited to damaged/depolarized mitochondria by PINK1, it ubiquitinates exposed mitochondrial proteins leading to both proteasomal degradation and mitophagy (Narendra et al, 2012; Sarraf et al, 2013). Parkin has also been implicated in cell surface signalling and as a tumour suppressor (see references in Wauer and Komander, 2013).Parkin encodes five structured domains, beginning with an N-terminal ubiquitin-like domain (UbLD) and followed by four domains that each bind two zinc (Zn) atoms (Figure 1A). The most N-terminal of the Zn-binding domains is RING0. C-terminal to this is the RBR, consisting of RING1, the IBR and RING2. The crystal structures of inactive Parkin from Riley et al (2013), Trempe et al (2013) and Wauer and Komander (2013) show remarkable congruity. Spatially, the IBR is at the complete opposite end of the molecule from RING2, to which it is connected by a partially unstructured ∼37 residue linker. This linker includes a two-turn helix, referred to as the repressor element of Parkin (REP) or tether, which binds and occludes the E2 binding face of RING1. RING1 occupies the central position in these structures, and RING0 separates RING1 from RING2 (Figure 1B and C). The latter contains the residue identified by Wenzel et al (2011), and confirmed by all three groups, to be the catalytic cysteine, C431. A lower resolution structure also includes the UbLD and places this domain adjacent to RING1 (Trempe et al, 2013). A second unstructured linker connects the UbLD and RING0. UbLDs are involved in a number of protein–protein interactions and small angle X-ray scattering confirms that this domain is integral to the core structure of Parkin (Spratt et al, 2013; Trempe et al, 2013). Biophysical characterization of Parkin and HHARI suggests that each is a monomer in solution.Open in a separate windowFigure 1Schematic and spatial representation of Parkin. (A) Primary structure and domain designations of Parkin, including the REP sequence within the otherwise unstructured IBR-RING2 linker. (B) Structural representation of full-length Parkin (PDB 4K95) highlighting the complex domain interactions in the three-dimensional structure, the catalytic C431 residue, and residue W403 within the REP, which plays a role in stabilizing the autoinhibited form of Parkin. (C) A model of Parkin with the E2 UbcH5B/Ube2D2 bound (devised using PDB 4K95 and PDB 4AP4 to mimic the position of an E2 bound to RING1) to illustrate the required displacement of UbLD and REP and the large distance between the E2∼Ub attachment site of the E2 and the catalytic active site of Parkin. Note that in this conformation the catalytic Cys within RING2 (C431) remains buried by RING0.RING1 is the only bona fide RING domain. All NMR and crystal structures of IBR domains from Parkin, HHARI and HOIP (PDB ID: 2CT7) are in good agreement. The Parkin and HHARI RING2s are structurally highly homologous and share a common Zn-coordinating arrangement with IBR domains. In contrast to the IBR and RING2, RING0 has a distinct arrangement of Zn-coordinating residues (Beasley et al, 2007; Duda et al, 2013; Riley et al, 2013; Spratt et al, 2013; Trempe et al, 2013; Wauer and Komander, 2013) (see Figure 1F of Trempe et al (2013) for the various Zn coordination arrangements).All of the Parkin crystal structures represent inactive forms of the E3. This is imposed by the quaternary positioning of the domains, which precludes activity in multiple ways. RING0 plays two obvious roles to maintain Parkin in an inactive state. RING0 shares an interface with RING2 and buries C431, making it unavailable as an ubiquitin acceptor. Moreover, RING0 intervenes between RING1 and RING2, creating an insurmountable separation of >50 Å between the active site Cys of an E2 bound to RING1 and C431 (Figure 1B and C). Thus, RING0 must be displaced for ubiquitin transfer to occur. Accordingly, deletion of RING0 results in a marked increase in Parkin autoubiquitination and in C431 reactivity (Riley et al, 2013; Trempe et al, 2013; Wauer and Komander, 2013). In HHARI, these two inhibitory functions are fulfilled by the C-terminal Ariadne domain, which similarly interposes between RING1 and RING2 (Duda et al, 2013).Additional inhibition is provided by the REP, which binds to RING1 at the canonical RING-E2 binding site and prevents E2 binding. This provides at least a partial explanation for the impaired ability of Parkin to bind E2 when compared to HHARI, which lacks this element (Duda et al, 2013). A disease-associated REP mutant (A398T) at the RING1 interface increases autoubiquitination (Wauer and Komander, 2013). The significance of inhibition by REP-RING1 binding was verified by mutating a critical RING1-interacting REP residue (W403A). This increased autoubiquitination and E2 binding (Trempe et al, 2013). Consistent with the requirement for charging C431 with ubiquitin in mitochondrial translocation (Lazarou et al, 2013), Parkin association with depolarized mitochondria is accelerated with this mutation (Trempe et al, 2013). Interestingly, W403 also interacts with the C-terminal Val of Parkin within RING2, and could therefore potentially further stabilize the autoinhibited form of the protein (Riley et al, 2013), consistent with previous observations (Henn et al, 2005).The quaternary structure of full-length Parkin also suggests that displacement of its N-terminal UbLD must occur for full activation (Trempe et al, 2013). The positioning of the UbLD adjacent to RING1 indicates that it would provide a steric impediment to E2∼Ub binding (Figure 1B and C). Additionally, displacement of the UbLD could be important to relieve interactions with the IBR-RING2 linker, which, as suggested in a previous study (Chaugule et al, 2011), might help to maintain Parkin in an inactive state. Finally, the crystal structure of the full-length Parkin indicates that the UbLD is not available for interactions with other proteins. This would limit Parkin''s range of intermolecular interactions.RBR E3s have at least two domains critical for sequential ubiquitin transfer and full activity, RING1 and RING2. The RING1 of Parkin, as well as all other RBR E3s, is notable in lacking the basic residue in the second Zn coordinating loop (or its equivalent in U-box proteins), which has recently been implicated in RING-mediated transfer of Ub from E2∼Ub (Metzger et al, 2013). This suggests that other factors play compensatory roles in positioning ubiquitin for transfer from E2∼Ub to C431. A non-mutually exclusive possibility is that the lack of this basic residue in RING1 limits unwanted attack on the E2∼Ub linkage, thereby minimizing the unregulated ubiquitination. Turning to RING2, the area surrounding the active site C431 of Parkin is notable in that it includes a sequence recognizable as a catalytic triad, similar to that in deubiquitinating enzymes. The Cys-His-Glu grouping, found in Parkin and other RBR E3s, contributes to in vitro activity (Riley et al, 2013; Wauer and Komander, 2013). Interestingly, however, the Glu was dispensable in a cellular assay (Riley et al, 2013). This triad is conserved in HHARI, where an Asn between the Cys and His residues (found in a number of RBRs but not conserved in Parkin), was found to be important for catalysis (Duda et al, 2013).The advances made in these studies impart significant information about an important and clinically relevant E3. However, Parkin, as well as HHARI, has been captured in their inactive, unmodified forms. One obvious question is how does Parkin transition between inactive and active states. PINK1 is implicated in phosphorylating Parkin on its UbLD and potentially other sites, with evidence that phosphorylation contributes to Parkin activation (Narendra et al, 2012). How phosphorylation could contribute to protein interactions that might facilitate Parkin activation, potentially including Parkin oligomerization (Lazarou et al, 2013), is unknown. Regardless, it is evident that considerable unwinding of its quaternary structure must take place.While there is much work ahead to understand these processes, one important interface that must be disrupted for activation is that between the REP and RING1. It is intriguing to consider that such interruption might be associated with other alterations in the IBR-RING2 linker, potentially facilitating the movement of the UbLD from RING1 and contributing to activation. Related to activation is the all-important question of how Parkin recognizes and targets specific substrates. While the UbLD represents a potential site of interaction, most purported substrates are not known to have UbLD-interaction domains. Although interactions involving the UbLD could occur indirectly, through bridging molecules, there is also evidence that other regions of Parkin, including the RBR region, might recognize substrates either directly or indirectly (Tsai et al, 2003) and that some substrates may be phosphorylated by PINK1 (Narendra et al, 2012). Conformational changes induced by substrate interactions, particularly in the IBR RING2 linker, could, as above, represent an important aspect of activation.There are over 75 missense mutations of Parkin associated with AR-JP, most of these inactivate the protein, but there are also some that are activating (Wauer and Komander, 2013). Activating mutations presumably result in pathology at least partially as a consequence of increased autoubiquitination and degradation (e.g., A398T). The current studies help to provide a classification of missense mutations into those that affect (i) folding or stability, (ii) catalytic mechanism, and (iii) interactions between domains. Interdomain mutations might inactivate or contribute to constitutive activation leading to autoubiquitination and degradation.Finally, we know little about how the autosomal recessive and the much more prevalent sporadic forms of PD overlap in their molecular pathology. However, mitochondrial dysfunction is increasingly a common theme. Thus, with the structure of the inactive protein in hand, there is hope that we can begin to consider ways in which domain interactions might be altered in a controlled manner to activate, but not hyperactivate, this critical E3 and lessen the progression of PD.  相似文献   

6.
7.
8.
Is there a cellular mechanism for preventing a depolymerizing microtubule track from “slipping out from under” its cargo? A recent study in budding yeast indicates that when a chromosome is transported to the minus end of a spindle microtubule, its kinetochore-bound microtubule plus end–tracking protein (+TIP) Stu2 may move to the plus end to promote rescue; i.e., to switch the depolymerizing end to a polymerizing end. The possibility that other +TIPs may play a similar role in sustaining a microtubule track during vesicular transport deserves investigation.Microtubule motor proteins such as dynein and kinesins are responsible for transporting cellular cargos along microtubule tracks (Vale, 2003). The net direction and speed of cargo movement, however, are likely to be regulated in a very complicated fashion, especially when a cargo is bound to multiple motors with opposite directionalities (Vale 2003; Mallik and Gross 2004; Levi et al., 2006). The fact that the microtubule track is not very stable further complicates matters. The plus ends of microtubules, which face the cell periphery in most cell types, are highly dynamic, exhibiting alternating periods of polymerization (growth) and depolymerization (shrinkage; Desai and Mitchison, 1997). Such plus end dynamics may be useful for searching and capturing relatively stationary cargos near the cell periphery that need to be transported inward (Vaughan et al., 2002). However, the dynamic nature of the track can also create an obvious problem for the transport process. If a microtubule''s rate of shrinkage is greater than the rate of cargo transport, then the microtubule may shrink past an attached cargo, causing its dissociation from the track. Does this happen in cells, or do cells have a mechanism to prevent it?Although this question has never been directly addressed, a recent study on budding yeast chromosome segregation has shed new light on the issue (Tanaka et al., 2005). In this study, the authors took advantage of a strategy that allowed them to specifically shut off the function of a single kinetochore, thereby preventing it from attaching to a spindle microtubule while, at the same time, permitting other kinetochores to attach to the spindle. After the function of this single kinetochore was switched back on, the behavior of its associated chromosome on a spindle microtubule was subjected to a detailed image analysis. Several important insights from this study on chromosome–microtubule interactions during mitosis have been recently reviewed (Bloom 2005), and, thus, only those observations that pertain to cargo transport will be highlighted here. The chromosome was first seen to undergo a lateral interaction with the microtubule followed by minus end–directed transport toward the pole. The mechanism of the minus end–directed transport is not entirely clear, although a member of the kinesin-14 family, Kar3, may be one of the players in this process (Tanaka et al., 2005). During transport, the attached microtubule can undergo shrinkage with a rate higher than that of the minus end–directed chromosome movement (Tanaka et al., 2005); however, it never shrank beyond the position of the cargo. Such exquisite control over the extent of shrinkage appears to rely on a conversation between the cargo and the plus end of the microtubule that is mediated by the microtubule plus end–tracking protein Stu2 (Bloom 2005; Tanaka et al., 2005).Microtubule plus end–tracking proteins (+TIPs) are a class of proteins that use different structural motifs or specific targeting mechanisms to localize to the dynamic plus ends of microtubules (Carvalho et al., 2003; Akhmanova and Hoogenraad, 2005). Although most +TIPs associate with only the growing ends of microtubules, several +TIPs also localize to the shrinking ends (Carvalho et al., 2003; 2004; Akhmanova and Hoogenraad, 2005; Mennella et al., 2005; Sproul et al., 2005; Molk et al., 2006; Wu et al., 2006). Many +TIPs have been found to impact microtubules by either promoting their growth or promoting dynamic behavior. Stu2 is a member of the XMAP215/TOG/Dis1/DdCP224 family of proteins that have been shown to affect microtubule dynamics in multiple ways depending on different experimental conditions (Ohkura et al., 2001; Popov and Karsenti 2003; Holmfeldt et al., 2004; Akhmanova and Hoogenraad, 2005). In vitro, Stu2 binds to the plus ends of preformed microtubules and promotes catastrophe, which is a switch from growth to shrinkage (van Breugel et al., 2003). In vivo studies using mutants of Stu2, however, indicate that Stu2 promotes microtubule growth (Severin et al., 2001) and the dynamics of both kinetochore and cytoplasmic microtubules (Kosco et al., 2001; Pearson et al., 2003). During anaphase B spindle elongation, Stu2 may antagonize the function of Kip3 (a kinesin-13 family member) to promote the plus end polymerization of overlapping microtubules (Severin et al., 2001). Although it is not fully understood how or why Stu2 is so versatile, it is well recognized that the in vivo interactions among +TIPs are very complicated, and the loss of function of a +TIP in vivo may decrease or increase the accumulation of other +TIPs that also regulate microtubule dynamics (Carvalho et al., 2003; 2004; Lansbergen et al., 2004; Akhmanova and Hoogenraad, 2005; Galjart 2005; Komarova et al., 2005). Tanaka et al. (2005) identified Stu2 as a rescue (a switch from shrinkage to growth) factor based not on phenotypic studies of Stu2 mutants but, instead, on a direct observation of the relationship between microtubule plus end behavior and Stu2 localization. They found that Stu2 was localized at the plus ends of microtubules emanating from the spindle pole body, and, during periods of microtubule shrinkage, Stu2 levels at the plus ends were decreased. Interestingly, Stu2 was also localized at the unbound kinetochore. When the kinetochore subsequently attached laterally to a spindle microtubule and underwent minus end–directed transport, the Stu2 proteins were transported from the kinetochore to the microtubule plus end. The arrival of Stu2 at the plus end closely correlated to the rescue of the shrinking microtubule (Tanaka et al., 2005). These observations strongly suggest that the Stu2 carried by the kinetochore may serve as a rescue factor for the microtubule track, preventing it from vanishing before the migrating chromosome.Could such a scenario exist during microtubule-dependent transport of nonchromosomal cargoes during interphase? We do not yet know the answer. However, based on published studies, it seems reasonable to hypothesize that other +TIPs, especially the cytoplasmic linker protein CLIP-170, may function in a manner similar to yeast Stu2 to ensure a safe trip for a minus end–directed cargo. CLIP-170 contains CAP-Gly microtubule-binding motifs at its NH2 terminus and was initially identified as a protein required for linking endocytic vesicles to microtubules in vitro (Pierre et al., 1992; Rickard and Kreis 1996). Later, CLIP-170 was identified as a founding member of the microtubule plus end–tracking proteins (Perez et al., 1999). The connection between CLIP-170''s in vitro endosome–microtubule linking property and its in vivo plus end tracking behavior has not been clearly made. Could an endocytic vesicle use its bound CLIP-170 as a rescue factor to prevent the disappearance of the track on which it is traveling?CLIP-170 is indeed considered to be a rescue factor in mammalian cells (Komarova et al., 2002a). Komarova et al. (2002b) have found that in cultured CHO and NRK cells, microtubule dynamics seem to be controlled spatially; catastrophe and rescue occur frequently only near the cell periphery. Although the mechanisms behind catastrophe and rescue are not fully understood, protein factors are required for regulating both events in vivo (Desai and Mitchison, 1997). In CHO cells, a dominant-negative form of CLIP-170 that displaces the endogenous CLIP-170 from microtubule plus ends severely reduces the rescue frequency so that microtubules are more likely to shrink all the way back to the microtubule-organizing center (Komarova et al., 2002a). Moreover, both in vivo and in vitro studies suggest that the rescue activity of CLIP-170 is localized to the NH2 terminus containing the CAP-Gly motifs (Komarova et al., 2002a; Arnal et al., 2004). How CLIP-170 rescues a shrinking end is not clear. CLIP-170 can promote tubulin oligomerization (Diamantopoulos et al., 1999; Arnal et al., 2004), and it is likely that this property serves to increase the local concentration of tubulin substrate, thereby lowering the entropic barrier for the polymerization reaction. CLIP-170 in mammalian cells has only been found at growing plus ends, most likely as a result of copolymerization with tubulin subunits followed by its release from older segments (Diamantopoulos et al., 1999; Perez et al., 1999; Folker et al., 2005). When a microtubule end shrinks, CLIP-170 falls off. Is there a mechanism to get CLIP-170 close to the depolymerizing end and facilitate its function as a rescue factor? Given the proposed function of Stu2 as a rescue factor for spindle microtubules, one may easily imagine a similar scenario in which vesicle-bound CLIP-170 may be transported to the approaching microtubule end to rescue it from further shrinkage.If vesicle-bound CLIP-170 is transported to the plus end in a manner similar to Stu2, could such transport be mediated by plus end–directed kinesins? Although the kinesin involved in transporting Stu2 toward the microtubule plus end still needs to be identified, detailed image analyses have revealed a role for the Kip2/Tea2 kinesins (members of the kinesin-7 family) in transporting CLIP-170 homologues in fungi (Busch et al., 2004; Carvalho et al., 2004). Bik1 and Tip1 are the CLIP-170 homologues in budding and fission yeasts, respectively, and these proteins are found at microtubule plus ends, where they act as growth-promoting factors or anticatastrophe factors (Berlin et al., 1990; Brunner and Nurse 2000; Carvalho et al., 2004). In both yeasts, the Kip2/Tea2 kinesins bind to and comigrate with the CLIP-170 homologues along the microtubule toward the plus end (Busch et al., 2004; Carvalho et al., 2004). Kinesins have also been implicated in targeting other +TIPs to microtubule plus ends (Jimbo et al., 2002; Maekawa et al., 2003; Zhang et al., 2003; Wu et al., 2006). For example, the mammalian tumor suppressor protein APC (adenomatous polyposis coli) may be targeted to the plus end by KIF3A/KIF3B (a heterotrimeric kinesin II in the kinesin-2 family) as well as by other mechanisms (Jimbo et al., 2002; Nathke 2004; Slep et al., 2005). It will be interesting to see whether a similar transport process for CLIP-170 exists in higher eukaryotic cells. It is possible that such a mechanism would deliver just enough CLIP-170 to the shrinking plus end to initiate rescue. When microtubule growth is resumed, CLIP-170''s intrinsic higher affinity for tubulin subunits and lower affinity for the microtubule wall may allow these proteins to “treadmill” on the growing end (Perez et al., 1999; Folker et al., 2005).The regulation of CLIP-170 activity appears to be rather complex. CLIP-170 is most likely phosphorylated by multiple kinases, including FKBP12–rapamycin-associated protein (mTOR; Choi et al., 2002). Although phosphorylation by mTOR/FKBP 12–rapamycin-associated protein may stimulate CLIP-170''s microtubule binding, phosphorylation by other kinases may cause CLIP-170 to dissociate from microtubules (Rickard and Kreis 1996; Choi et al., 2002). In vivo, CLIP-170 has a closed conformation that is presumably inactive and an open conformation that may interact with microtubules and dynein regulators such as dynactin (Schroer 2004) and LIS1 (Morris et al., 1998; Lansbergen et al., 2004). It is possible that phosphorylation may regulate the conversion between these two forms, but the specific mechanism and the spatial regulation for this conversion have yet to be resolved. If CLIP-170 is indeed released from a membranous cargo to move to the plus end in order to serve as a rescue factor, it would be interesting to know when and/or where such a conformational switch occurs. Finally, other proteins may play redundant roles with CLIP-170 in vesicular trafficking, which may explain why a dramatic defect in vesicle/organelle distribution is not detected when the CLIP-170 level is lowered or when the gene is knocked out (Lansbergen et al., 2004; Akhmanova et al., 2005).+TIPs other than CLIP-170 may play a similar role in rescuing shrinking microtubule tracks. For example, the dynactin complex that links dynein to membranous cargoes and promotes the processive motion of dynein (Schroer 2004) may act as a rescue factor. The p150Glued subunit of dynactin and CLIP-170 both contain CAP-Gly microtubule-binding motifs at their NH2 termini, although p150Glued contains one, whereas CLIP-170 contains two such motifs. Dynactin has been shown to behave as a +TIP facilitating the capture of vesicular cargo for minus end–directed transport (Vaughan et al., 1999; 2002). The head domain of the p150Glued subunit containing the CAP-Gly motif has been shown to promote rescue in vivo in the absence of endogenous CLIP-170, although the effect was much weaker than that caused by the exogenous CLIP-170 head domain (Kamarova et al., 2002a). In vitro studies showed that dynactin may promote nucleation during microtubule assembly (Ligon et al., 2003), which is consistent with it being a potential rescue factor. As shown with CLIP-170, this capacity to bring multiple tubulins together may help to overcome the entropic barrier of the polymerization reaction. Finally, cargo-bound dynactin may also use kinesin to get to the plus end. The p150Glued subunit of dynactin has been shown to interact directly with the COOH terminus of KAP3, a subunit of the heterotrimeric kinesin II (a member of the kinesin-2 family) that also binds to APC (Jimbo et al., 2002; Deacon et al., 2003; Dell 2003). Although this binding is implicated in dynactin''s role as a cargo adaptor for kinesin II, it is possible, in theory, that a small amount of dynactin may use this connection to move to the plus end.The proposed hypothesis that +TIPs may be released from a membranous cargo to rescue a shrinking microtubule track may apply to both minus and plus end–directed transport. In addition, it is important to point out that this hypothesis does not exclude other mechanisms for rescuing long microtubule tracks. Rescue may occur stochastically, and, sometimes, +TIPs may participate in other ways such as mediating microtubule capture by the actin-rich cortex to stabilize the track (Wen et al., 2004; Galjart 2005). In some situations, microtubule dynamics are modulated by the direct binding of membranous cargo to the growing or shrinking plus ends of microtubules (Waterman-Storer and Salmon, 1998).Currently, the ability of CLIP-170 or other +TIPs to be released from a membranous cargo and to act as a rescue factor for a shrinking microtubule is just a hypothesis. Nevertheless, searching for proteins involved in the communication between a cargo and the approaching shrinking end of its microtubule track is clearly an endeavor worth pursuing.  相似文献   

9.
EMBO J (2013) 32 23, 3029–3040 10.1038/emboj.2013.223; published online October112013Primary cilia are cell surface sensory organelles, whose dysfunction underlies various human genetic diseases collectively termed ciliopathies. A new study in The EMBO Journal by Villumsen et al now reveals how stress–response pathways converge to stimulate ciliogenesis by modulating protein composition of centriolar satellites. Better understanding of these mechanisms should bring us closer to identifying the cellular defects that underlie ciliopathies caused by mutations in centriolar satellite proteins.Centrioles are barrel-shaped structures with two distinct identities. In proliferating cells centrioles provide structural support for the centrosome, a key microtubule-organizing centre, whereas in quiescent cells centrioles are converted into basal bodies and promote the assembly of primary cilia. In centrosomes, centrioles are embedded in pericentriolar material (PCM), a dynamic structure responsible for microtubule nucleation. PCM proteins exhibit cell cycle-dependent localisation, achieved at least in part by the regulation of their transport. Centriolar satellites, dense fibrous granules frequently clustered around the interphase centrosome, have been implicated in microtubule-dependent protein transport to centrosomes (Kubo et al, 1999). In particular, PCM-1, the core constituent of centriolar satellites, is required for centrosomal accumulation of several PCM components (Dammermann and Merdes, 2002). Although the proteomic composition of satellites is still elusive, the growing list of satellite proteins includes CEP131/AZI1 (Staples et al, 2012), CEP290 (Stowe et al, 2012), Bardet-Biedl syndrome protein 4 (BBS4) and Oral facial digital syndrome protein (OFD1; Lopes et al, 2011). Mutations in OFD1, CEP290 and BBS4 cause ciliopathies (Kim et al, 2008), underscoring a functional link between satellites and ciliogenesis. So far, two roles have been proposed for satellites in cilia formation: First, in cycling cells they may serve to sequester essential ciliary proteins (Stowe et al, 2012). Second, upon initiation of the ciliogenesis programme, centriolar satellite components seem to promote the recruitment of specific ciliary proteins to basal bodies (Ferrante et al, 2006; Lopes et al, 2011; Stowe et al, 2012).In a new study in The EMBO Journal, Villumsen et al (2013) now describe how stress–response pathways conspire to control ciliogenesis. The authors observed that specific environmental stresses, such as ultraviolet light radiation (UV) or heat shock, but not ionizing radiation (IR), trigger rapid displacement of PCM-1, AZI1 and CEP290 from centriolar satellites. However, OFD1 remained associated with satellites, indicating that centriolar satellites persist despite UV-induced removal of PCM-1. This might come as some surprise, since PCM-1 depletion by RNA interference (RNAi) is thought to disrupt satellite integrity (Kim et al, 2008; Lopes et al, 2011); however, satellite loss upon PCM-1 RNAi may be a consequence of prolonged depletion of PCM-1, while acute PCM-1 displacement by stress might only ‘remodel'' centriolar satellites. It is also possible that not all satellites are created equal, and they do vary in protein composition (Kim et al, 2008; Staples et al, 2012). If so, UV-induced PCM-1 removal may disrupt some, but not all satellites.A good candidate regulator of centriolar satellite remodelling was the stress-activated MAP kinase p38, and indeed, Villumsen et al (2013) found p38 MAPK activity to be stimulated by both UV and heat shock but not IR in U2OS cells, mirroring those very stress pathways that also cause displacement of AZI1 and PCM-1 from satellites. Furthermore, p38 MAPK was essential for UV-induced dispersal of PCM-1 and AZI1. The authors then tested the hypothesis that stress-induced centriolar satellite remodelling could involve changes in the interactome of AZI1, and—consistent with an earlier proteomics study (Akimov et al, 2011)—identified PCM-1, CEP290 and the mindbomb E3 ubiquitin protein ligase 1 (MIB1) as the main AZI1 binding partners. GFP-MIB1 localized to centriolar satellites and mono-ubiquitylated AZI1, PCM-1 and CEP290 in cycling cells. In response to UV, both ubiquitylation of these proteins and MIB1 activity were reduced; notably, UV-induced MIB1 inactivation was independent of p38 MAPK activity, indicating that these two enzymes may act via distinct pathways (Figure 1A).Open in a separate windowFigure 1(A) Regulation of centriolar satellite remodelling. (B) Schematic summary of how centriolar satellite remodelling might facilitate ciliogenesis. See text for details.What could be the purpose of MIB1-dependent ubiquitylation of these satellite proteins? It certainly does not seem to regulate subcellular targeting, as in MIB1-depleted cells, AZI1 and PCM-1 both localised normally to centriolar satellites and could still be displaced by UV. Instead, ubiquitylation seems to suppress the interaction between AZI1 and PCM-1, consistent with the observation that UV, a condition that also reduces their ubiquitylation, enhances the binding of AZI1 to PCM-1.PCM-1, CEP290 and AZI1 all participate in ciliogenesis (Kim et al, 2008; Wilkinson et al, 2009; Stowe et al, 2012), raising the possibility that MIB1 might also affect this process. Indeed, serum starvation, which is known to promote cilia formation, attenuated MIB1 activity. Furthermore, MIB1 overexpression reduced the ciliogenesis observed in serum-starved cells, while MIB1 depletion in proliferating cells triggered a marked increase in the proportion of cells that formed cilia; this seems to reflect a direct effect of MIB1 on ciliogenesis, since neither MIB1 depletion nor overexpression altered cell cycle progression. Taken together, downregulation of MIB1 enzymatic activity appears to be a pre-requisite for efficient ciliogenesis, regardless of whether it is triggered by physiological ciliogenesis-promoting signals or by environmental stresses, making MIB1 a novel negative regulator of cilia formation.The recent discovery of ciliopathy-associated mutations in constituents of the DNA damage response signalling pathway pointed to a connection between DNA damage and ciliogenesis (Chaki et al, 2012). With the new link between UV and centriolar satellites, the authors next asked if UV radiation might affect ciliogenesis. Remarkably, UV and heat shock both triggered cilia assembly in RPE-1 cells in a p38 MAPK-dependent manner. MIB1 depletion further enhanced ciliogenesis after UV radiation, again implying an additive effect of p38 MAPK signalling and MIB1 suppression (Figure 1A).While finer details on the precise role of centriolar satellite components in cilia formation are still lacking, a more coherent picture is finally starting to emerge. In cycling cells, ubiquitination by MIB1 could serve to limit the interaction between AZI1 and PCM-1 on centriolar satellites (Figure 1B). Under these conditions PCM-1 may bind and sequester CEP290, an essential ciliogenic protein, thereby precluding untimely cilia formation (Stowe et al, 2012). Both during normal and stress-induced ciliogenesis programs, remodelling of centriolar satellites creates a permissive environment for cilia formation, and a key step in this process is downregulation of MIB1 activity. While it remains to be established how the latter is achieved, it is clear that MIB1 inactivation causes loss of ubiquitylation and increased binding between AZI1 and PCM-1. Preferential interaction of PCM-1 with AZI1 could in turn facilitate release of CEP290 from centriolar satellites and its subsequent accumulation at the centrosome. Once CEP290 reaches the optimum concentration at the centriole/basal body, it could serve to tether AZI1–PCM-1 complexes. PCM-1 could then concentrate Rab8 GTPase near centrosomes, allowing CEP290 to recruit Rab8 into the cilium, where it acts to extend the ciliary membrane (Kim et al, 2008).Collectively, the findings reported here provide strong experimental support to the notion that centriolar satellites are negative regulators of ciliogenesis in proliferating cells. Their role is central to limit untimely formation of cilia in cells. Environmental strains elicit stress–response pathways that converge to relieve the ciliogenesis block imposed by satellites. It is tempting to speculate that stress-induced cilia might serve as signalling platforms and contribute to checkpoint activation or perhaps initiation of repair mechanisms, but more work is needed to establish the true purpose of ciliogenesis in this context. It is of considerable interest that a recent study reports that autophagy, another stress-induced pathway, selectively removes OFD1 from satellites to promote ciliogenesis (Tang et al, 2013). Therefore stress-mediated centriolar satellite remodelling seems to be an evolving theme in the control of ciliogenesis.  相似文献   

10.
Microtubules are cytoskeletal filaments that are dynamically assembled from α/β-tubulin heterodimers. The primary sequence and structure of the tubulin proteins and, consequently, the properties and architecture of microtubules are highly conserved in eukaryotes. Despite this conservation, tubulin is subject to heterogeneity that is generated in two ways: by the expression of different tubulin isotypes and by posttranslational modifications (PTMs). Identifying the mechanisms that generate and control tubulin heterogeneity and how this heterogeneity affects microtubule function are long-standing goals in the field. Recent work on tubulin PTMs has shed light on how these modifications could contribute to a “tubulin code” that coordinates the complex functions of microtubules in cells.

Introduction

Microtubules are key elements of the eukaryotic cytoskeleton that dynamically assemble from heterodimers of α- and β-tubulin. The structure of microtubules, as well as the protein sequences of α- and β-tubulin, is highly conserved in evolution, and consequently, microtubules look alike in almost all species. Despite the high level of conservation, microtubules adapt to a large variety of cellular functions. This adaptation can be mediated by a large panel of microtubule-associated proteins (MAPs), including molecular motors, as well as by mechanisms that directly modify the microtubules, thus either changing their biophysical properties or attracting subsets of MAPs that convey specific functions to the modified microtubules. Two different mechanism can generate microtubule diversity: the expression of different α- and β-tubulin genes, referred to as tubulin isotypes, and the generation of posttranslational modifications (PTMs) on α- and β-tubulin (Figs. 1 and and2).2). Although known for several decades, deciphering how tubulin heterogeneity controls microtubule functions is still largely unchartered. This review summarizes the current advances in the field and discusses new concepts arising.Open in a separate windowFigure 1.Tubulin heterogeneity generated by PTMs. (A) Schematic representation of the distribution of different PTMs of tubulin on the α/β-tubulin dimer with respect to their position in the microtubule lattice. Acetylation (Ac), phosphorylation (P), and polyamination (Am) are found within the tubulin bodies that assemble into the microtubule lattice, whereas polyglutamylation, polyglycylation, detyrosination, and C-terminal deglutamylation take place within the C-terminal tubulin tails that project away from the lattice surface. The tubulin dimer represents TubA1A and TubB2B (Fig. 2), and modification sites for polyglutamylation and polyglycylation have been randomly chosen. (B) Chemical structure of the branched peptide formed by polyglutamylation and polyglycylation, using the γ-carboxyl groups of the modified glutamate residues as acceptor sites for the isopeptide bonds. Note that in the case of polyglutamylation, the elongation of the side chains generates classical peptide bonds (Redeker et al., 1991).Open in a separate windowFigure 2.Heterogeneity of C-terminal tails of tubulin isotypes and their PTMs. The amino acid sequences of all tubulin genes found in the human genome are indicated, starting at the last amino acid of the folded tubulin bodies. Amino acids are represented in single-letter codes and color coded according to their biochemical properties. Known sites for polyglutamylation are indicated (Eddé et al., 1990; Alexander et al., 1991; Rüdiger et al., 1992). Potential modification sites (all glutamate residues) are indicated. Known C-terminal truncation reactions of α/β-tubulin (tub) are indicated. The C-terminal tails of the yeast Saccharomyces cerevisiae are shown to illustrate the phylogenetic diversity of these domains.

Tubulin isotypes

The cloning of the first tubulin genes in the late 1970’s (Cleveland et al., 1978) revealed the existence of multiple genes coding for α- or β-tubulin (Ludueña and Banerjee, 2008) that generate subtle differences in their amino acid sequences, particularly in the C-terminal tails (Fig. 2). It was assumed that tubulin isotypes, as they were named, assemble into discrete microtubule species that carry out unique functions. This conclusion was reinforced by the observation that some isotypes are specifically expressed in specialized cells and tissues and that isotype expression changes during development (Lewis et al., 1985; Denoulet et al., 1986). These high expectations were mitigated by a subsequent study showing that all tubulin isotypes freely copolymerize into heterogeneous microtubules (Lewis et al., 1987). To date, only highly specialized microtubules, such as ciliary axonemes (Renthal et al., 1993; Raff et al., 2008), neuronal microtubules (Denoulet et al., 1986; Joshi and Cleveland, 1989), and microtubules of the marginal band of platelets (Wang et al., 1986; Schwer et al., 2001) are known to depend on some specific (β) tubulin isotypes, whereas the function of most other microtubules appears to be independent of their isotype composition.More recently, a large number of mutations in single tubulin isotypes have been linked to deleterious neurodevelopmental disorders (Keays et al., 2007; Fallet-Bianco et al., 2008; Tischfield et al., 2010; Cederquist et al., 2012; Niwa et al., 2013). Mutations of a single tubulin isotype could lead to an imbalance in the levels of tubulins as a result of a lack of incorporation of mutant isoforms into the microtubule lattice or to incorporation that perturbs the architecture or dynamics of the microtubules. The analysis of tubulin disease mutations is starting to reveal how subtle alterations of the microtubule cytoskeleton can lead to functional aberrations in cells and organisms and might provide novel insights into the roles of tubulin isotypes that have so far been considered redundant.

Tubulin PTMs

Tubulin is subject to a large range of PTMs (Fig. 1), from well-known ones, such as acetylation or phosphorylation, to others that have so far mostly been found on tubulin. Detyrosination/tyrosination, polyglutamylation, and polyglycylation, for instance, might have evolved to specifically regulate tubulin and microtubule functions, in particular in cilia and flagella, as their evolution is closely linked to these organelles. The strong link between those modifications and tubulin evolution has led to the perception that they are tubulin PTMs; however, apart from detyrosination/tyrosination, most of them have other substrates (Regnard et al., 2000; Xie et al., 2007; van Dijk et al., 2008; Rogowski et al., 2009).

Tubulin acetylation.

Tubulin acetylation was discovered on lysine 40 (K40; Fig. 1 A) of flagellar α-tubulin in Chlamydomonas reinhardtii (L’Hernault and Rosenbaum, 1985) and is generally enriched on stable microtubules in cells. Considering that K40 acetylation per se has no effect on the ultrastructure of microtubules (Howes et al., 2014), it is rather unlikely that it directly stabilizes microtubules. As a result of its localization at the inner face of microtubules (Soppina et al., 2012), K40 acetylation might rather affect the binding of microtubule inner proteins, a poorly characterized family of proteins (Nicastro et al., 2011; Linck et al., 2014). Functional experiments in cells have further suggested that K40 acetylation regulates intracellular transport by regulating the traffic of kinesin motors (Reed et al., 2006; Dompierre et al., 2007). These observations could so far not be confirmed by biophysical measurements in vitro (Walter et al., 2012; Kaul et al., 2014), suggesting that in cells, K40 acetylation might affect intracellular traffic by indirect mechanisms.Enzymes involved in K40 acetylation are HDAC6 (histone deacetylase family member 6; Hubbert et al., 2002) and Sirt2 (sirtuin type 2; North et al., 2003). Initial functional studies used overexpression, depletion, or chemical inhibition of these enzymes. These studies should be discussed with care, as both HDAC6 and Sirt2 deacetylate other substrates and have deacetylase-independent functions and chemical inhibition of HDAC6 is not entirely selective for this enzyme (Valenzuela-Fernández et al., 2008). In contrast, acetyl transferase α-Tat1 (or Mec-17; Akella et al., 2010; Shida et al., 2010) specifically acetylates α-tubulin K40 (Fig. 3), thus providing a more specific tool to investigate the functions of K40 acetylation. Knockout mice of α-Tat1 are completely void of K40-acetylated tubulin; however, they show only slight phenotypic aberrations, for instance, in their sperm flagellum (Kalebic et al., 2013). A more detailed analysis of α-Tat1 knockout mice demonstrated that absence of K40 acetylation leads to reduced contact inhibition in proliferating cells (Aguilar et al., 2014). In migrating cells, α-Tat1 is targeted to microtubules at the leading edge by clathrin-coated pits, resulting in locally restricted acetylation of those microtubules (Montagnac et al., 2013). A recent structural study of α-Tat1 demonstrated that the low catalytic rate of this enzyme, together with its localization inside the microtubules, caused acetylation to accumulate selectively in stable, long-lived microtubules (Szyk et al., 2014), thus explaining the link between this PTM and stable microtubules in cells. However, the direct cellular function of K40 acetylation on microtubules is still unclear.Open in a separate windowFigure 3.Enzymes involved in PTM of tubulin. Schematic representation of known enzymes (mammalian enzymes are shown) involved in the generation and removal of PTMs shown in Fig. 1. Note that some enzymes still remain unknown, and some modifications are irreversible. (*CCP5 preferentially removes branching points [Rogowski et al., 2010]; however, the enzyme can also hydrolyze linear glutamate chains [Berezniuk et al., 2013]).Recent discoveries have brought up the possibility that tubulin could be subject to multiple acetylation events. A whole-acetylome study identified >10 novel sites on α- and β-tubulin (Choudhary et al., 2009); however, none of these sites have been confirmed. Another acetylation event has been described at lysine 252 (K252) of β-tubulin. This modification is catalyzed by the acetyltransferase San (Fig. 3) and might regulate the assembly efficiency of microtubules as a result of its localization at the polymerization interface (Chu et al., 2011).

Tubulin detyrosination.

Most α-tubulin genes in different species encode a C-terminal tyrosine residue (Fig. 2; Valenzuela et al., 1981). This tyrosine can be enzymatically removed (Hallak et al., 1977) and religated (Fig. 3; Arce et al., 1975). Mapping of tyrosinated and detyrosinated microtubules in cells using specific antibodies (Gundersen et al., 1984; Geuens et al., 1986; Cambray-Deakin and Burgoyne, 1987a) revealed that subsets of interphase and mitotic spindle microtubules are detyrosinated (Gundersen and Bulinski, 1986). As detyrosination was mostly found on stable and long-lived microtubules, especially in neurons (Cambray-Deakin and Burgoyne, 1987b; Robson and Burgoyne, 1989; Brown et al., 1993), it was assumed that this modification promotes microtubule stability (Gundersen et al., 1987; Sherwin et al., 1987). Although a direct stabilization of the microtubule lattice was considered unlikely (Khawaja et al., 1988), it was found more recently that detyrosination protects cellular microtubules from the depolymerizing activity of kinesin-13–type motor proteins, such as KIF2 or MCAK, thus increasing their longevity (Peris et al., 2009; Sirajuddin et al., 2014).Besides kinesin-13 motors, plus end–tracking proteins with cytoskeleton-associated protein glycine-rich (CAP-Gly) domains, such as CLIP170 or p150/glued, specifically interact with tyrosinated microtubules (Peris et al., 2006; Bieling et al., 2008) via this domain (Honnappa et al., 2006). In contrast, kinesin-1 moves preferentially on detyrosinated microtubules tracks in cells (Liao and Gundersen, 1998; Kreitzer et al., 1999; Konishi and Setou, 2009). The effect of detyrosination on kinesin-1 motor behavior was recently measured in vitro, and a small but significant increase in the landing rate and processivity of the motor has been found (Kaul et al., 2014). Such subtle changes in the motor behavior could, in conjunction with other factors, such as regulatory MAPs associated with cargo transport complexes (Barlan et al., 2013), lead to a preferential use of detyrosinated microtubules by kinesin-1 in cells.Despite the early biochemical characterization of a detyrosinating activity, the carboxypeptidase catalyzing detyrosination of α-tubulin has yet to be identified (Hallak et al., 1977; Argaraña et al., 1978, 1980). In contrast, the reverse enzyme, tubulin tyrosine ligase (TTL; Fig. 3; Raybin and Flavin, 1975; Deanin and Gordon, 1976; Argaraña et al., 1980), has been purified (Schröder et al., 1985) and cloned (Ersfeld et al., 1993). TTL modifies nonpolymerized tubulin dimers exclusively. This selectivity is determined by the binding interface between the TTL and tubulin dimers (Szyk et al., 2011, 2013; Prota et al., 2013). In contrast, the so far unidentified detyrosinase acts preferentially on polymerized microtubules (Kumar and Flavin, 1981; Arce and Barra, 1983), thus modifying a select population of microtubules within cells (Gundersen et al., 1987).In most organisms, only one unique gene for TTL exists. Consequently, TTL knockout mice show a huge accumulation of detyrosinated and particularly Δ2-tubulin (see next section). TTL knockout mice die before birth (Erck et al., 2005) with major developmental defects in the nervous system that might be related to aberrant neuronal differentiation (Marcos et al., 2009). TTL is strictly tubulin specific (Prota et al., 2013), indicating that all observed defects in TTL knockout mice are directly related to the deregulation of the microtubule cytoskeleton.

Δ2-tubulin and further C-terminal modification.

A biochemical study of brain tubulin revealed that ∼35% of α-tubulin cannot be retyrosinated (Paturle et al., 1989) because of the lack of the penultimate C-terminal glutamate residue of the primary protein sequence (Fig. 2; Paturle-Lafanechère et al., 1991). This so-called Δ2-tubulin (for two C-terminal amino acids missing) cannot undergo retyrosination as a result of structural constraints within TTL (Prota et al., 2013) and thus is considered an irreversible PTM.Δ2-tubulin accumulates in long-lived microtubules of differentiated neurons, axonemes of cilia and flagella, and also in cellular microtubules that have been artificially stabilized, for instance, with taxol (Paturle-Lafanechère et al., 1994). The generation of Δ2-tubulin requires previous detyrosination of α-tubulin; thus, the levels of this PTM are indirectly regulated by the detyrosination/retyrosination cycle. This mechanistic link is particularly apparent in the TTL knockout mice, which show massive accumulation of Δ2-tubulin in all tested tissues (Erck et al., 2005). Loss of TTL and the subsequent increase of Δ2-tubulin levels were also linked to tumor growth and might contribute to the aggressiveness of the tumors by an as-yet-unknown mechanism (Lafanechère et al., 1998; Mialhe et al., 2001). To date, no specific biochemical role of Δ2-tubulin has been determined; thus, one possibility is that the modification simply locks tubulin in the detyrosinated state.The enzymes responsible for Δ2-tubulin generation are members of a family of cytosolic carboxypeptidases (CCPs; Fig. 3; Kalinina et al., 2007; Rodriguez de la Vega et al., 2007), and most of them also remove polyglutamylation from tubulin (see next section; Rogowski et al., 2010). These enzymes are also able to generate Δ3-tubulin (Fig. 1 A; Berezniuk et al., 2012), indicating that further degradation of the tubulin C-terminal tails are possible; however, the functional significance of this event is unknown.

Polyglutamylation.

Polyglutamylation is a PTM that occurs when secondary glutamate side chains are formed on γ-carboxyl groups of glutamate residues in a protein (Fig. 1, A and B). The modification was first discovered on α- and β-tubulin from the brain (Eddé et al., 1990; Alexander et al., 1991; Rüdiger et al., 1992; Mary et al., 1994) as well as on axonemal tubulin from different species (Mary et al., 1996, 1997); however, it is not restricted to tubulin (Regnard et al., 2000; van Dijk et al., 2008). Using a glutamylation-specific antibody, GT335 (Wolff et al., 1992), it was observed that tubulin glutamylation increases during neuronal differentiation (Audebert et al., 1993, 1994) and that axonemes of cilia and flagella (Fouquet et al., 1994), as well as centrioles of mammalian centrosomes (Bobinnec et al., 1998), are extensively glutamylated.Enzymes catalyzing polyglutamylation belong to the TTL-like (TTLL) family (Regnard et al., 2003; Janke et al., 2005). In mammals, nine glutamylases exist, each of them showing intrinsic preferences for modifying either α- or β-tubulin as well as for initiating or elongating glutamate chains (Fig. 3; van Dijk et al., 2007). Two of the six well-characterized TTLL glutamylases also modify nontubulin substrates (van Dijk et al., 2008).Knockout or depletion of glutamylating enzymes in different model organisms revealed an evolutionarily conserved role of glutamylation in cilia and flagella. In motile cilia, glutamylation regulates beating behavior (Janke et al., 2005; Pathak et al., 2007; Ikegami et al., 2010) via the regulation of flagellar dynein motors (Kubo et al., 2010; Suryavanshi et al., 2010). Despite the expression of multiple glutamylases in ciliated cells and tissues, depletion or knockout of single enzymes often lead to ciliary defects, particularly in motile cilia (Ikegami et al., 2010; Vogel et al., 2010; Bosch Grau et al., 2013; Lee et al., 2013), suggesting essential and nonredundant regulatory functions of these enzymes in cilia.Despite the enrichment of polyglutamylation in neuronal microtubules (Audebert et al., 1993, 1994), knockout of TTLL1, the major polyglutamylase in brain (Janke et al., 2005), did not show obvious neuronal defects in mice (Ikegami et al., 2010; Vogel et al., 2010). This suggests a tolerance of neuronal microtubules to variations in polyglutamylation.Deglutamylases, the enzymes that reverse polyglutamylation, were identified within a novel family of CCPs (Kimura et al., 2010; Rogowski et al., 2010). So far, three out of six mammalian CCPs have been shown to cleave C-terminal glutamate residues, thus catalyzing both the reversal of polyglutamylation and the removal of gene-encoded glutamates from the C termini of proteins (Fig. 3). The hydrolysis of gene-encoded glutamate residues is not restricted to tubulin, in which it generates Δ2- and Δ3-tubulin, but has also been reported for other proteins such as myosin light chain kinase (Rusconi et al., 1997; Rogowski et al., 2010). One enzyme of the CCP family, CCP5, preferentially removes branching points generated by glutamylation, thus allowing the complete reversal of the polyglutamylation modification (Kimura et al., 2010; Rogowski et al., 2010). However, CCP5 can also hydrolyze C-terminal glutamate residues from linear peptide chains similar to other members of the CCP family (Berezniuk et al., 2013).CCP1 is mutated in a well-established mouse model for neurodegeneration, the pcd (Purkinje cell degeneration) mouse (Mullen et al., 1976; Greer and Shepherd, 1982; Fernandez-Gonzalez et al., 2002). The absence of a key deglutamylase leads to strong hyperglutamylation in brain regions that undergo degeneration, such as the cerebellum and the olfactory bulb (Rogowski et al., 2010). When glutamylation levels were rebalanced by depletion or knockout of the major brain polyglutamylase TTLL1 (Rogowski et al., 2010; Berezniuk et al., 2012), Purkinje cells survived. Although the molecular mechanisms of hyperglutamylation-induced degeneration remain to be elucidated, perturbation of neuronal transport, as well as changes in the dynamics and stability of microtubules, is expected to be induced by hyperglutamylation. Increased polyglutamylation levels have been shown to affect kinesin-1–mediated transport in cultured neurons (Maas et al., 2009), and the turnover of microtubules can also be regulated by polyglutamylation via the activation of microtubule-severing enzymes such as spastin (Lacroix et al., 2010).Subtle differences in polyglutamylation can be seen on diverse microtubules in different cell types. The functions of these modifications remain to be studied; however, its wide distribution strengthens the idea that it could be involved in fine-tuning a range of microtubule functions.

Polyglycylation.

Tubulin polyglycylation or glycylation, like polyglutamylation, generates side chains of glycine residues within the C-terminal tails of α- and β-tubulin (Fig. 1, A and B). The modification sites of glycylation are considered to be principally the same as for glutamylation, and indeed, both PTMs have been shown to be interdependent in cells (Rogowski et al., 2009; Wloga et al., 2009). Initially discovered on Paramecium tetraurelia tubulin (Redeker et al., 1994), glycylation has been extensively studied using two antibodies, TAP952 and AXO49 (Bressac et al., 1995; Levilliers et al., 1995; Bré et al., 1996). In contrast to polyglutamylation, glycylation is restricted to cilia and flagella in most organisms analyzed so far.Glycylating enzymes are also members of the TTLL family, and homologues of these enzymes have so far been found in all organisms with proven glycylation of ciliary axonemes (Rogowski et al., 2009; Wloga et al., 2009). In mammals, initiating (TTLL3 and TTLL8) and elongating (TTLL10) glycylases work together to generate polyglycylation (Fig. 3). In contrast, the two TTLL3 orthologues from Drosophila melanogaster can both initiate and elongate glycine side chains (Rogowski et al., 2009).In mice, motile ependymal cilia in brain ventricles acquire monoglycylation upon maturation, whereas polyglycylation is observed only after several weeks (Bosch Grau et al., 2013). Sperm flagella, in contrast, acquire long glycine chains much faster, suggesting that the extent of polyglycylation could correlate with the length of the axonemes (Rogowski et al., 2009). Depletion of glycylases in mice (ependymal cilia; Bosch Grau et al., 2013), zebrafish (Wloga et al., 2009; Pathak et al., 2011), Tetrahymena thermophila (Wloga et al., 2009), and D. melanogaster (Rogowski et al., 2009) consistently led to ciliary disassembly or severe ciliary defects. How glycylation regulates microtubule functions remains unknown; however, the observation that glycylation-depleted axonemes disassemble after initial assembly (Rogowski et al., 2009; Bosch Grau et al., 2013) suggests a role of this PTM in stabilizing axonemal microtubules. Strikingly, human TTLL10 is enzymatically inactive; thus, humans have lost the ability to elongate glycine side chains (Rogowski et al., 2009). This suggests that the elongation of the glycine side chains is not an essential aspect of the function of this otherwise critical tubulin PTM.

Other tubulin PTMs.

Several other PTMs have been found on tubulin. Early studies identified tubulin phosphorylation (Eipper, 1974; Gard and Kirschner, 1985; Díaz-Nido et al., 1990); however, no specific functions were found. The perhaps best-studied phosphorylation event on tubulin takes place at serine S172 of β-tubulin (Fig. 1 A), is catalyzed by the Cdk1 (Fig. 3), and might regulate microtubule dynamics during cell division (Fourest-Lieuvin et al., 2006; Caudron et al., 2010). Tubulin can be also modified by the spleen tyrosine kinase Syk (Fig. 3; Peters et al., 1996), which might play a role in immune cells (Faruki et al., 2000; Sulimenko et al., 2006) and cell division (Zyss et al., 2005; Sulimenko et al., 2006).Polyamination has recently been discovered on brain tubulin (Song et al., 2013), after having been overlooked for many years as a result of the low solubility of polyaminated tubulin. Among several glutamine residues of α- and β-tubulin that can be polyaminated, Q15 of β-tubulin is considered the primary modification site (Fig. 1 A). Polyamination is catalyzed by transglutaminases (Fig. 3), which modify free tubulin as well as microtubules in an irreversible manner, and most likely contribute to the stabilization of microtubules (Song et al., 2013).Tubulin was also reported to be palmitoylated (Caron, 1997; Ozols and Caron, 1997; Caron et al., 2001), ubiquitinated (Ren et al., 2003; Huang et al., 2009; Xu et al., 2010), glycosylated (Walgren et al., 2003; Ji et al., 2011), arginylated (Wong et al., 2007), methylated (Xiao et al., 2010), and sumoylated (Rosas-Acosta et al., 2005). These PTMs have mostly been reported without follow-up studies, and some of them are only found in specific cell types or organisms and/or under specific metabolic conditions. Further studies will be necessary to gain insights into their potential roles for the regulation of the microtubule cytoskeleton.

Current advances and future perspectives

The molecular heterogeneity of microtubules, generated by the expression of different tubulin isotypes and by the PTM of tubulin has fascinated the scientific community for ∼40 years. Although many important advances have been made in the past decade, the dissection of the molecular mechanisms and a comprehensive understanding of the biological functions of tubulin isotypes and PTMs will be a challenging field of research in the near future.

Direct measurements of the impact of tubulin heterogeneity.

The most direct and reliable type of experiments to determine the impact of tubulin heterogeneity on microtubule behavior are in vitro measurements with purified proteins. However, most biophysical work on microtubules has been performed with tubulin purified from bovine, ovine, or porcine brains, which can be obtained in large quantities and with a high degree of purity and activity (Vallee, 1986; Castoldi and Popov, 2003). Brain tubulin is a mixture of different tubulin isotypes and is heavily posttranslationally modified and thus inept for investigating the functions of tubulin heterogeneity (Denoulet et al., 1986; Cambray-Deakin and Burgoyne, 1987b; Paturle et al., 1989; Eddé et al., 1990). Thus, pure, recombinant tubulin will be essential to dissect the roles of different tubulin isoforms and PTMs.Attempts to produce recombinant, functional α- and β-tubulin in bacteria have failed so far (Yaffe et al., 1988), most likely because of the absence of the extensive tubulin-specific folding machinery (Yaffe et al., 1992; Gao et al., 1993; Tian et al., 1996; Vainberg et al., 1998) in prokaryotes. An alternative source of tubulin with less isotype heterogeneity and with almost no PTMs is endogenous tubulin from cell lines such as HeLa, which in the past has been purified using a range of biochemical procedures (Bulinski and Borisy, 1979; Weatherbee et al., 1980; Farrell, 1982; Newton et al., 2002; Fourest-Lieuvin, 2006). Such tubulin can be further modified with tubulin-modifying enzymes, such as polyglutamylases, either by expressing those enzymes in the cells before tubulin purification (Lacroix and Janke, 2011) or in vitro with purified enzymes (Vemu et al., 2014). Despite some technical limitations of these methods, HeLa tubulin modified in cells has been successfully used in an in vitro study on the role of polyglutamylation in microtubule severing (Lacroix et al., 2010).Naturally occurring variants of tubulin isotypes and PTMs can be purified from different organisms, organs, or cell types, but obviously, only some combinations of tubulin isotypes and PTMs can be obtained by this approach. The recent development of an affinity purification method using the microtubule-binding TOG (tumor overexpressed gene) domain of yeast Stu2p has brought a new twist to this approach, as it allows purifying small amounts of tubulin from any cell type or tissue (Widlund et al., 2012).The absence of tubulin heterogeneity in yeast has made budding and fission yeast potential expression systems for recombinant, PTM-free tubulin (Katsuki et al., 2009; Drummond et al., 2011; Johnson et al., 2011). However, the expression of mammalian tubulin in this system has remained impossible. This problem was then partially circumvented by expressing tubulin chimeras that consist of a yeast tubulin body fused to mammalian C-terminal tubulin tails, thus mimicking different tubulin isotypes (Sirajuddin et al., 2014). Moreover, detyrosination can be generated by deleting the key C-terminal residue from endogenous or chimeric α-tubulin (Badin-Larçon et al., 2004), and polyglutamylation is generated by chemically coupling glutamate side chains to specifically engineered tubulin chimeras (Sirajuddin et al., 2014). These approaches allowed the first direct measurements of the impact of tubulin isotypes and PTMs on the behavior of molecular motors in vitro (Sirajuddin et al., 2014) and the analysis of the effects of tubulin heterogeneity on microtubule behavior and interactions inside the yeast cell (Badin-Larçon et al., 2004; Aiken et al., 2014).Currently, the most promising development has been the successful purification of fully functional recombinant tubulin from the baculovirus expression system (Minoura et al., 2013). Using this system, defined α/β-tubulin dimers can be obtained using two different epitope tags on α- and β-tubulin, respectively. Although these epitope tags are essential for separating recombinant from the endogenous tubulin, they could also affect tubulin assembly or microtubule–MAP interactions. Thus, future developments should focus on eliminating these tags.Current efforts have brought the possibility of producing recombinant tubulin into reach. Further improvement and standardization of these methods will certainly provide a breakthrough in understanding the mechanisms by which tubulin heterogeneity contributes to microtubule functions.

Complexity of tubulin—understanding the regulatory principles.

The diversity of tubulin genes (isotypes) and the complexity of tubulin PTMs have led to the proposal of the term “tubulin code” (Verhey and Gaertig, 2007; Wehenkel and Janke, 2014), in analogy to the previously coined histone code (Jenuwein and Allis, 2001). Tubulin molecules consist of a highly structured and thus evolutionarily conserved tubulin body and the unstructured and less conserved C-terminal tails (Nogales et al., 1998). As PTMs and sequence variations within the tubulin body are expected to affect the conserved tubulin fold and therefore the properties of the microtubule lattice, they are not likely to be involved in generating the tubulin code. In contrast, modulations of the C-terminal tails could encode signals on the microtubule surface without perturbing basic microtubule functions and properties (Figs. 1 A and and4).4). Indeed, the highest degree of gene-encoded diversity (Fig. 2) and the highest density and complexity of PTMs (Fig. 1) are found within these tail domains.Open in a separate windowFigure 4.Molecular components of the tubulin code. Schematic representation of potential coding elements that could generate specific signals for the tubulin code. (A) The length of the C-terminal tails of different tubulin isotypes differ significantly (Fig. 2) and could have an impact on the interactions between microtubules and MAPs. (B) Tubulin C-terminal tails are rich in charged amino acid residues. The distribution of these residues and local densities of charges could influence the electrostatic interactions with the tails and the readers. (C) Although each glutamate residue within the C-terminal tails could be considered a potential modification site, only some sites have been found highly occupied in tubulin purifications from native sources. This indicates selectivity of the modification reactions, which can participate in the generation of specific modification patterns (see D). Modification sites might be distinguished by their neighboring amino acid residues, which could create specific modification epitopes. (D) As a result of the large number of modification sites and the variability of side chains, a large variety of modification patterns could be generated within a single C-terminal tail of tubulin. (E) Modification patterns as shown in D can be distinct between α- and β-tubulin. These modification patterns could be differentially distributed at the surface of the microtubule lattice, thus generating a higher-order patterning. Tub, tubulin. For color coding, see Fig. 2.Considering the number of tubulin isotypes plus all potential combinations of PTMs (e.g., each glutamate residue within the C-terminal tubulin tail could be modified by either polyglutamylation or polyglycylation, each of them generating side chains of different lengths; Fig. 4), the number of distinct signals generated by the potential tubulin code would be huge. However, as many of these potential signals represent chemical structures that are similar and might not be reliably distinguished by readout mechanisms, it is possible that the tubulin code generates probabilistic signals. In this scenario, biochemically similar modifications would have similar functional readouts, and marginal differences between those signals would only bias biological processes but not determine them. This stands in contrast to the concept of the histone code, in which precise patterns of different PTMs on the histone proteins encode distinct biological signals.The concept of probabilistic signaling is already inscribed in the machinery that generates the tubulin code. Polyglutamylases and polyglycylases from the TTLL family have preferential activities for either α- or β-tubulin and for generating different lengths of the branched glutamate or glycine chains. Although under conditions of low enzyme concentrations, as found in most cells and tissues, the enzymes seem to selectively generate their preferential type of PTM, higher enzyme concentrations induce a more promiscuous behavior, leading, for instance, to a loss of selectivity for α- or β-tubulin (van Dijk et al., 2007). Similarly, the modifying enzymes might prefer certain modification sites within the C-terminal tails of tubulin but might be equally able to modify other sites, which could be locally regulated in cells. For example, β-tubulin isotypes isolated from mammalian brain were initially found to be glutamylated on single residues (Alexander et al., 1991; Rüdiger et al., 1992), which in the light of the comparably low sensitivity of mass spectrometry at the time might rather indicate a preferential than a unique modification of these sites. Nevertheless, the neuron-specific polyglutamylase for β-tubulin TTLL7 (Ikegami et al., 2006) can incorporate glutamate onto many more modification sites of β-tubulin in vitro (Mukai et al., 2009), which clearly indicates that not all of the possible modification events take place under physiological conditions.Several examples supporting a probabilistic signaling mode of the tubulin code are found in the recent literature. In T. thermophila, a ciliate without tubulin isotype diversity (Gaertig et al., 1993) but with a huge repertoire of tubulin PTMs and tubulin-modifying enzymes (Janke et al., 2005), tubulin can be easily mutagenized to experimentally eliminate sites for PTMs. Mutagenesis of the most commonly occupied glutamylation/glycylation sites within the β-tubulin tails did not generate a clear decrease of glycylation levels nor did it cause obvious phenotypic alterations. This indicates that the modifying enzymes can deviate toward alternative modification sites and that similar PTMs on different sites can compensate the functions of the mutated site. However, when all of the key modification sites were mutated, glycylation became prominently decreased, which led to severe phenotypes, including lethality (Xia et al., 2000). Most strikingly, these phenotypes could be recovered by replacing the C-terminal tail of α-tubulin with the nonmutated β-tubulin tail. This α–β-tubulin chimera became overglycylated and functionally compensated for the absence of modification sites on β-tubulin. The conclusion of this study is that PTM- and isotype-generated signals can fulfill a biological function within a certain range of tolerance.But how efficient is such compensation? The answer can be found in a variety of already described deletion mutants for tubulin-modifying enzymes in different model organisms. Most single-gene knockouts for TTLL genes (glutamylases or glycylases) did not result in prominent phenotypic alterations in mice, even for enzymes that are ubiquitously expressed. Only some highly specialized microtubule structures show functional aberrations upon the deletion of a single enzyme. These “tips of the iceberg” are usually the motile cilia and sperm flagella, which carry very high levels of polyglutamylation and polyglycylation (Bré et al., 1996; Kann et al., 1998; Rogowski et al., 2009). It thus appears that some microtubules are essentially dependent on the generation of specific PTM patterns, whereas others can tolerate changes and appear to function normally. How “normal” these functions are remains to be investigated in future studies. It is possible that defects are subtle and thus overlooked but could become functionally important under specific conditions.A tubulin code also requires readout mechanisms. The most likely “readers” of the tubulin code are MAPs and molecular motors. Considering the probabilistic signaling hypothesis, the expected effects of the signals would be in most cases rather gradual changes, for instance, to fine-tune molecular motor traffic and/or to bias motors toward defined microtubule tracks but not to obliterate motor activity or MAP binding to microtubules. An in vitro study using recombinant tubulin chimeras purified from yeast confirmed this notion (Sirajuddin et al., 2014). By analyzing which elements of the tubulin code can regulate the velocity and processivity of the molecular motors kinesin and dynein, these researchers found that the C-terminal tails of α- and β-tubulin differentially influence the kinetic parameters of the tested motors; however, the modulation was rather modest. One of their striking observations was that a single lysine residue, present in the C-terminal tails of two β-tubulin isotypes (Figs. 2 and and4),4), significantly affected motor traffic and that this effect can be counterbalanced by polyglutamylation. These observations are the first in vitro evidence for the interdependence of different elements of the tubulin code and provide another indication for its probabilistic mode of signaling.

Future directions.

One of the greatest technological challenges to understanding the function of the tubulin code is to detect and interpret subtle and complex regulatory events generated by this code. It will thus be instrumental to further develop tools to better distinguish graded changes in PTM levels on microtubules in cells and tissues (Magiera and Janke, 2013) and to reliably measure subtle modulations of microtubule behavior in reconstituted systems.The current advances in the field and especially the availability of whole-organism models, as well as first insights into the pathological role of tubulin mutations (Tischfield et al., 2011), are about to transform our way of thinking about the regulation of microtubule cytoskeleton. Tubulin heterogeneity generates complex probabilistic signals that cannot be clearly attributed to single biological functions in most cases and that are not essential for most cellular processes. Nevertheless, it has been conserved throughout evolution of eukaryotes and can hardly be dismissed as not important. To understand the functional implications of these processes, we might be forced to reconsider how we define biologically important events and how we measure events that might encode probabilistic signals. The answers to these questions could provide novel insights into how complex systems, such as cells and organisms, are sustained throughout difficult and challenging life cycles, resist to environmental stress and diseases, and have the flexibility needed to succeed in evolution.  相似文献   

11.
12.
13.
14.
15.
To gain a further understanding of axonemal dynein regulation, mutant strains of Chlamydomonas reinhardtii that had defects in both phototactic behavior and flagellar motility were identified and characterized. ptm1, ptm2, and ptm3 mutant strains exhibited motility phenotypes that resembled those of known inner dynein arm region mutant strains, but did not have biochemical or genetic phenotypes characteristic of other inner dynein arm mutations. Three other mutant strains had defects in the f class of inner dynein arms. Dynein extracts from the pf9-4 strain were missing the entire f complex. Strains with mutations in pf9/ida1, ida2, or ida3 failed to assemble the f dynein complex and did not exhibit phototactic behavior. Fractionated dynein from mia1-1 and mia2-1 axonemes exhibited a novel f class inner dynein arm biochemical phenotype; the 138-kD f intermediate chain was present in altered phosphorylation forms. In vitro axonemal dynein activity was reduced by the mia1-1 and mia2-1 mutations. The addition of kinase inhibitor restored axonemal dynein activity concomitant with the dephosphorylation of the 138-kD f intermediate chain. Dynein extracts from uni1-1 axonemes, which specifically assemble only one of the two flagella, contained relatively high levels of the altered phosphorylation forms of the 138-kD intermediate chain. We suggest that the f dynein complex may be phosphoregulated asymmetrically between the two flagella to achieve phototactic turning. C hlamydomonas reinhardtii flagella use an asymmetric beat stroke, similar to a breast stroke, to propel cells forward. To generate the asymmetric beat stroke, dynein activity must be regulated both along the length and around the circumference of the flagella. If all dyneins were active at the same time, the flagella would exist in a state of rigor. The dyneins are located in two rows along the length of the doublet microtubules. The inner dynein arms are heterogeneous in composition with at least eight heavy chains and various intermediate and light chains arranged in an elaborate morphology that repeats every 96 nm (Kagami and Kamiya, 1992; Mastronarde et al., 1992). In contrast, the outer dynein arms are biochemically and morphologically homogeneous (Huang et al., 1979; Mitchell and Rosenbaum, 1985; Kamiya, 1988); each outer dynein arm contains three dynein heavy chains and 10 intermediate and light chains. The inner and outer arms appear to have different functions in the formation of the beat stroke; the inner arms generate the waveform of the beat stroke, whereas the outer arms provide additional force to the waveform (Brokaw and Kamiya, 1987).Previous workers had shown that dynein regulation is imposed, in part, by activities of the radial spokes and the central pair complex. Mutant strains that are missing or have altered radial spokes or central pair complexes are paralyzed even if they have a full complement of dyneins (Adams et al., 1981; Piperno et al., 1981). Many extragenic suppressors of this paralysis phenotype do not restore the missing structures, but rather suppress by altering either inner arm or outer arm region structures (Huang et al., 1982a ; Piperno et al., 1992; Porter et al., 1992, 1994). These data suggest that direct or indirect interactions exist between the dynein arms and the radial spokes or central pair complexes.Over 80 proteins in Chlamydomonas flagella are phosphorylated (Piperno et al., 1981), which makes dynein regulation by phosphorylation an attractive model. Hasegawa et al. (1987) showed that a higher percentage of demembranated axonemes reactivate with ATP after treatments that lower cAMP levels or inhibit cAMP-dependent protein kinase (cAPK)1. In flagella from other organisms, cAMP has an opposite role (for reviews see Tash and Means, 1983; Tash, 1989). An increased frequency of reactivation also occurs after the NP-40–soluble components are extracted from the axonemes, which suggests that the cAPK, target phosphoproteins, and endogenous phosphatases are all integral axonemal components (Hasegawa et al., 1987). In quantitative sliding disintegration assays, the inner dynein arm activity of axonemes that are missing the radial spokes is increased in the presence of pharmacological or specific peptide inhibitors of cAPK (Smith and Sale, 1992; Howard et al., 1994). Reconstitution experiments with axonemes that are missing the radial spokes suggest that radial spokes normally function to activate the inner dynein arms by inhibiting a cAPK (Smith and Sale, 1992; Howard et al., 1994). It is not known if the cAPK directly phosphorylates inner dynein arm components or phosphorylates another axonemal component that then acts on the inner dynein arms (Howard et al., 1994).The f (originally called I1) inner arms are biochemically the best studied inner dynein arm complex. This complex is comprised of two dynein heavy chains and three intermediate chains of 140, 138, and 110 kD; it can be purified by sucrose density centrifugation (Piperno and Luck, 1981; Smith and Sale, 1991; Porter et al., 1992) or ion-exchange chromatography (Kagami and Kamiya, 1992). The purified complex has low ATPase activity and only rarely translocates microtubules in vitro (Smith and Sale, 1991; Kagami and Kamiya, 1992). Deep-etch EM of the purified f inner arm shows a two-headed complex that is connected to a common base by thin stalks (Smith and Sale, 1991). Longitudinal EM image analyses have shown that this complex is located just proximally of the first radial spoke in each 96-nm repeating unit (Piperno et al., 1990; Mastronarde et al., 1992). Mutations at three different loci (PF9/ IDA1, IDA2, and IDA3) result in the complete loss of the f complex (Kamiya et al., 1991; Kagami and Kamiya, 1992; Porter et al., 1992). The PF9/IDA1 locus encodes a dynein heavy chain that is believed to be one of the two heavy chains that are components of the f complex (Porter, 1996).We undertook a new approach to identify axonemal components involved in dynein regulation; we isolated and characterized mutant strains that were unable to perform phototaxis. In Chlamydomonas, phototaxis is a behavior by which cells orient to the direction of incident light. Light direction is detected by the eyespot, an asymmetrically located organelle, and a signal is transmitted to the flagella using voltage-gated ion channels (Harz and Hegemann, 1991). For cells to perform phototaxis, the waveforms of the two flagella are altered coordinately. The trans flagellum, which is located farther from the eyespot, beats with a larger front amplitude than the cis flagellum to turn the cell toward the light (Rüffer and Nultsch, 1991). It seemed likely that the alterations in the beat amplitudes needed for correct phototactic behavior could be caused by differential dynein regulation in the cis and trans flagella. Therefore, we hypothesized that there should be a class of phototactic mutant strains that is not able to perform phototaxis because of defects in the regulation of dyneins. Three of the eight phototactic mutant strains that we characterized had biochemical defects in the f class of inner dynein arms. One of these strains, pf9-4, was missing the entire f complex, and the other two strains, mia1-1 and mia2-1, exhibited a novel f class inner dynein arm biochemical phenotype. These observations suggest that the f inner dynein arm is a target for regulation during phototaxis.  相似文献   

16.
17.
Nek2 (for NIMA-related kinase 2) is a mammalian cell cycle–regulated kinase structurally related to the mitotic regulator NIMA of Aspergillus nidulans. In human cells, Nek2 associates with centrosomes, and overexpression of active Nek2 has drastic consequences for centrosome structure. Here, we describe the molecular characterization of a novel human centrosomal protein, C-Nap1 (for centrosomal Nek2-associated protein 1), first identified as a Nek2-interacting protein in a yeast two-hybrid screen. Antibodies raised against recombinant C-Nap1 produced strong labeling of centrosomes by immunofluorescence, and immunoelectron microscopy revealed that C-Nap1 is associated specifically with the proximal ends of both mother and daughter centrioles. On Western blots, anti–C-Nap1 antibodies recognized a large protein (>250 kD) that was highly enriched in centrosome preparations. Sequencing of overlapping cDNAs showed that C-Nap1 has a calculated molecular mass of 281 kD and comprises extended domains of predicted coiled-coil structure. Whereas C-Nap1 was concentrated at centrosomes in all interphase cells, immunoreactivity at mitotic spindle poles was strongly diminished. Finally, the COOH-terminal domain of C-Nap1 could readily be phosphorylated by Nek2 in vitro, as well as after coexpression of the two proteins in vivo. Based on these findings, we propose a model implicating both Nek2 and C-Nap1 in the regulation of centriole–centriole cohesion during the cell cycle.The serine/threonine kinase NIMA of Aspergillus nidulans is considered the founding member of a family of protein kinases with a possible role in cell cycle regulation (for reviews see Fry and Nigg, 1995; Lu and Hunter, 1995a ; Osmani and Ye, 1996). In A. nidulans, NIMA clearly cooperates with the Cdc2 protein kinase to promote progression into mitosis (Osmani et al., 1991), and overexpression of NIMA in a variety of heterologous species promotes a premature onset of chromosome condensation (O''Connell et al., 1994; Lu and Hunter, 1995b ). This has been interpreted to suggest evolutionary conservation of a pathway involving NIMA-related kinases (for review see Lu and Hunter, 1995a ). Indeed, kinases structurally related to NIMA are present in many species (Fry and Nigg, 1997). However, the only bona fide functional homologue of NIMA so far isolated stems from another filamentous fungus, Neurospora crassa (Pu et al., 1995), and the functional relationship between vertebrate NIMA-related kinases and fungal NIMA remains uncertain.The closest known mammalian relative to NIMA is a kinase termed Nek2 (for NIMA-related kinase 2)1 (Fry and Nigg, 1997). This kinase undergoes cell cycle–dependent changes in abundance and activity, reminiscent of NIMA (Schultz et al., 1994; Fry et al., 1995). It is highly expressed in male germ cells (Rhee and Wolgemuth, 1997; Tanaka et al., 1997), and data have been reported consistent with a role for Nek2 in meiotic chromosome condensation (Rhee and Wolgemuth, 1997). However, overexpression of active Nek2 in somatic cells has no obvious effect on chromosome condensation; instead, it induces striking alterations in the structure of the centrosome, the principal microtubule-organizing center of mammalian cells (Fry et al., 1998). Furthermore, immunofluorescence microscopy and subcellular fractionation concur to demonstrate that endogenous Nek2 associates with centrosomes, strongly suggesting that one physiological function of this kinase may relate to the centrosome cycle (Fry et al., 1998).The mammalian centrosome is an organelle of about 1 μm in diameter. It comprises two barrel-shaped centrioles that are made of nine short triplet microtubules and are surrounded by an amorphous matrix known as the pericentriolar material (PCM) (for review see Brinkley, 1985; Vorobjev and Nadehzdina, 1987; Kimble and Kuriyama, 1992; Kalt and Schliwa, 1993; Kellogg et al., 1994; Lange and Gull, 1996). Major progress has recently been made with the demonstration that microtubules are nucleated from γ-tubulin–containing ring complexes (γ-TuRCs), which are concentrated within the PCM (Moritz et al., 1995; Zheng et al., 1995). γ-Tubulin forms complexes with Spc97/98, two evolutionarily conserved proteins first identified in budding yeast spindle pole bodies (Geissler et al., 1996; Knop et al., 1997; Stearns and Winey, 1997), and there is also evidence for an important role of pericentrin and other coiled-coil proteins in organizing γ-TuRCs into higher order lattice structures (Doxsey et al., 1994; Dictenberg et al., 1998). However, in spite of this recent progress, it is clear that the inventory of centrosome components is far from complete.Centrosome structure and function is regulated in a cell cycle–dependent manner (for reviews see Mazia, 1987; Kellogg et al., 1994; Tournier and Bornens, 1994). Once in every cell cycle, and beginning around the G1/S transition, centrioles are duplicated (e.g., Kuriyama and Borisy, 1981a ; Vorobjev and Chentsov, 1982; Kochanski and Borisy, 1990; Chrétien et al., 1997). Late in G2, centrosomes then grow in size (a process referred to as maturation) through the recruitment of additional PCM proteins (Rieder and Borisy, 1982; Kalt and Schliwa, 1993; Lange and Gull, 1995). At the G2/M transition, the duplicated centrosomes separate and migrate to opposite ends of the nucleus. Concomitantly, their microtubule-nucleating activities increase dramatically in preparation for spindle formation (McGill and Brinkley, 1975; Snyder and McIntosh, 1975; Gould and Borisy, 1977; Kuriyama and Borisy, 1981b ; for reviews see Brinkley, 1985; Vorobjev and Nadehzdina, 1987; Karsenti, 1991). By what mechanisms these events are controlled remains largely unknown, but data obtained using phosphoepitope-specific antibodies strongly suggest that phosphorylation of centrosomal proteins plays a major role (Vandré et al., 1984, 1986; Centonze and Borisy, 1990). More direct support for this view stems from the observation that cyclin-dependent kinases (CDKs) enhance the microtubule-nucleation activity of centrosomes at the G2/M transition (Verde et al., 1990, 1992; Buendia et al., 1992) and are involved in promoting centrosome separation (Blangy et al., 1995; Sawin and Mitchison, 1995). Similarly, polo-like kinase 1, a cell cycle regulatory kinase structurally distinct from CDKs, has recently been implicated in centrosome maturation (Lane and Nigg, 1996).The precise role of Nek2 at the centrosome remains to be determined, but it is intriguing that overexpression of this kinase in human cells causes a pronounced splitting of centrosomes. This led us to propose that Nek2-dependent phosphorylation of previously unidentified proteins may cause a loss of centriole–centriole cohesion, and that this event might represent an early step in centrosome separation at the G2/M transition (Fry et al., 1998). With the aim of identifying potential substrates (or regulators) of Nek2, we have now performed a yeast two-hybrid screen, using full-length Nek2 as a bait. We report here the molecular characterization of a novel coiled-coil protein that we call C-Nap1 (for centrosomal Nek2-associated protein 1). C-Nap1 represents a core component of the mammalian centrosome and the first candidate substrate for a member of the NIMA protein kinase family to be identified.  相似文献   

18.
The erythropoietin receptor (EpoR) was discovered and described in red blood cells (RBCs), stimulating its proliferation and survival. The target in humans for EpoR agonists drugs appears clear—to treat anemia. However, there is evidence of the pleitropic actions of erythropoietin (Epo). For that reason, rhEpo therapy was suggested as a reliable approach for treating a broad range of pathologies, including heart and cardiovascular diseases, neurodegenerative disorders (Parkinson’s and Alzheimer’s disease), spinal cord injury, stroke, diabetic retinopathy and rare diseases (Friedreich ataxia). Unfortunately, the side effects of rhEpo are also evident. A new generation of nonhematopoietic EpoR agonists drugs (asialoEpo, Cepo and ARA 290) have been investigated and further developed. These EpoR agonists, without the erythropoietic activity of Epo, while preserving its tissue-protective properties, will provide better outcomes in ongoing clinical trials. Nonhematopoietic EpoR agonists represent safer and more effective surrogates for the treatment of several diseases such as brain and peripheral nerve injury, diabetic complications, renal ischemia, rare diseases, myocardial infarction, chronic heart disease and others.In principle, the erythropoietin receptor (EpoR) was discovered and described in red blood cell (RBC) progenitors, stimulating its proliferation and survival. Erythropoietin (Epo) is mainly synthesized in fetal liver and adult kidneys (13). Therefore, it was hypothesized that Epo act exclusively on erythroid progenitor cells. Accordingly, the target in humans for EpoR agonists drugs (such as recombinant erythropoietin [rhEpo], in general, called erythropoiesis-simulating agents) appears clear (that is, to treat anemia). However, evidence of a kaleidoscope of pleitropic actions of Epo has been provided (4,5). The Epo/EpoR axis research involved an initial journey from laboratory basic research to clinical therapeutics. However, as a consequence of clinical observations, basic research on Epo/EpoR comes back to expand its clinical therapeutic applicability.Although kidney and liver have long been considered the major sources of synthesis, Epo mRNA expression has also been detected in the brain (neurons and glial cells), lung, heart, bone marrow, spleen, hair follicles, reproductive tract and osteoblasts (617). Accordingly, EpoR was detected in other cells, such as neurons, astrocytes, microglia, immune cells, cancer cell lines, endothelial cells, bone marrow stromal cells and cells of heart, reproductive system, gastrointestinal tract, kidney, pancreas and skeletal muscle (1827). Conversely, Sinclair et al.(28) reported data questioning the presence or function of EpoR on nonhematopoietic cells (endothelial, neuronal and cardiac cells), suggesting that further studies are needed to confirm the diversity of EpoR. Elliott et al.(29) also showed that EpoR is virtually undetectable in human renal cells and other tissues with no detectable EpoR on cell surfaces. These results have raised doubts about the preclinical basis for studies exploring pleiotropic actions of rhEpo (30).For the above-mentioned data, a return to basic research studies has become necessary, and many studies in animal models have been initiated or have already been performed. The effect of rhEpo administration on angiogenesis, myogenesis, shift in muscle fiber types and oxidative enzyme activities in skeletal muscle (4,31), cardiac muscle mitochondrial biogenesis (32), cognitive effects (31), antiapoptotic and antiinflammatory actions (3337) and plasma glucose concentrations (38) has been extensively studied. Neuro- and cardioprotection properties have been mainly described. Accordingly, rhEpo therapy was suggested as a reliable approach for treating a broad range of pathologies, including heart and cardiovascular diseases, neurodegenerative disorders (Parkinson’s and Alzheimer’s disease), spinal cord injury, stroke, diabetic retinopathy and rare diseases (Friedreich ataxia).Unfortunately, the side effects of rhEpo are also evident. Epo is involved in regulating tumor angiogenesis (39) and probably in the survival and growth of tumor cells (25,40,41). rhEpo administration also induces serious side effects such as hypertension, polycythemia, myocardial infarction, stroke and seizures, platelet activation and increased thromboembolic risk, and immunogenicity (4246), with the most common being hypertension (47,48). A new generation of nonhematopoietic EpoR agonists drugs have hence been investigated and further developed in animals models. These compounds, namely asialoerythropoietin (asialoEpo) and carbamylated Epo (Cepo), were developed for preserving tissue-protective properties but reducing the erythropoietic activity of native Epo (49,50). These drugs will provide better outcome in ongoing clinical trials. The advantage of using nonhematopoietic Epo analogs is to avoid the stimulation of hematopoiesis and thereby the prevention of an increased hematocrit with a subsequent procoagulant status or increased blood pressure. In this regard, a new study by van Rijt et al. has shed new light on this topic (51). A new nonhematopoietic EpoR agonist analog named ARA 290 has been developed, promising cytoprotective capacities to prevent renal ischemia/reperfusion injury (51). ARA 290 is a short peptide that has shown no safety concerns in preclinical and human studies. In addition, ARA 290 has proven efficacious in cardiac disorders (52,53), neuropathic pain (54) and sarcoidosis-induced chronic neuropathic pain (55). Thus, ARA 290 is a novel nonhematopoietic EpoR agonist with promising therapeutic options in treating a wide range of pathologies and without increased risks of cardiovascular events.Overall, this new generation of EpoR agonists without the erythropoietic activity of Epo while preserving tissue-protective properties of Epo will provide better outcomes in ongoing clinical trials (49,50). Nonhematopoietic EpoR agonists represent safer and more effective surrogates for the treatment of several diseases, such as brain and peripheral nerve injury, diabetic complications, renal ischemia, rare diseases, myocardial infarction, chronic heart disease and others.  相似文献   

19.
The gene VI product (P6) of Cauliflower mosaic virus (CaMV) is a multifunctional protein known to be a major component of cytoplasmic inclusion bodies formed during CaMV infection. Although these inclusions are known to contain virions and are thought to be sites of translation from the CaMV 35S polycistronic RNA intermediate, the precise role of these bodies in the CaMV infection cycle remains unclear. Here, we examine the functionality and intracellular location of a fusion between P6 and GFP (P6-GFP). We initially show that the ability of P6-GFP to transactivate translation is comparable to unmodified P6. Consequently, our work has direct application for the large body of literature in which P6 has been expressed ectopically and its functions characterized. We subsequently found that P6-GFP forms highly motile cytoplasmic inclusion bodies and revealed through fluorescence colocalization studies that these P6-GFP bodies associate with the actin/endoplasmic reticulum network as well as microtubules. We demonstrate that while P6-GFP inclusions traffic along microfilaments, those associated with microtubules appear stationary. Additionally, inhibitor studies reveal that the intracellular movement of P6-GFP inclusions is sensitive to the actin inhibitor, latrunculin B, which also inhibits the formation of local lesions by CaMV in Nicotiana edwardsonii leaves. The motility of P6 along microfilaments represents an entirely new property for this protein, and these results imply a role for P6 in intracellular and cell-to-cell movement of CaMV.Cauliflower mosaic virus (CaMV), the type member of the genus Caulimovirus, has a circular double-stranded DNA genome known to encode six open reading frames (ORFs). The gene product of ORF VI (P6) is a multifunctional protein whose ascribed functions have increased in number since its initial characterization over 20 years ago. P6 was originally described as the most abundant CaMV protein in infected plants (Odell and Howell, 1980) and was later shown to be the major constituent of amorphous, electron-dense inclusion bodies that are thought to be the sites of virion assembly (Fujisawa et al., 1967; Rubio-Huertos et al., 1968; Himmelbach et al., 1996; Cecchini et al., 1997). Indeed, despite the detection of other viral proteins in CaMV inclusions, the P6 protein on its own is capable of forming inclusion bodies (Cecchini et al., 1997; Li and Leisner, 2002; Haas et al., 2005).P6 is the major pathogenicity determinant for CaMV (Daubert et al., 1984; Baughman et al., 1988; Stratford and Covey, 1989; Zijlstra and Hohn, 1992) and was recently shown to be a suppressor of RNA silencing (Love et al., 2007). In addition, P6 also functions as an avirulence determinant, as it has been shown to be responsible for eliciting a hypersensitive response in Nicotiana edwardsonii and Datura stramonium, as well as nonnecrotic resistance in Nicotiana bigelovii and Arabidopsis (Arabidopsis thaliana) ectotype Tsu-O (Daubert et al., 1984; Schoelz et al., 1986; Wintermantel et al., 1993; Agama et al., 2002). The portion of the P6 protein recognized by plants is localized to the N-terminal third of the protein (Wintermantel et al., 1993; Palanichelvam et al., 2000; Agama et al., 2002). P6 also has a significant effect on plant metabolism, as it is responsible for down-regulating or inducing expression of several plant genes (Geri et al., 1999), including genes involved in ethylene signaling (Geri et al., 2004).Replication of CaMV involves the production of a polycistronic RNA intermediate, the 35S RNA, and P6 acts as a translational transactivator (TAV) by modifying the host translational machinery to allow for reinitiation of translation on this RNA (Ryabova et al., 2002). To carry out this function, the P6 protein physically interacts with the initiation factor eIF3 (Park et al., 2001), as well as ribosomal proteins L13 (Bureau et al., 2004), L18 (Leh et al., 2000), and L24 (Park et al., 2001). Finally, P6 is also a nucleocytoplasmic shuttle protein whose nuclear export is dependent upon a Leu-rich sequence near its N terminus, a region that is also involved in inclusion body formation (Li and Leisner, 2002; Haas et al., 2005). Although the precise role of the P6 protein''s nucleocytoplasmic shuttle function during infection remains to be elucidated, P6 does have the capacity to bind RNA (De Tapia et al., 1993; Cerritelli et al., 1998) and as such may act to control export of the 35S RNA from the nucleus to the cytoplasm, drawing the 35S RNA into the nascent P6 inclusion bodies where viral proteins are translated.Despite the recognized intracellular movement of P6 from cytoplasm to nucleus and the disparate cytoplasmic functions of this protein, factors controlling intracellular transport of P6 remain unknown. The cytoskeleton has been implicated in the intracellular trafficking of a number of plant viral proteins. For example, proteins encoded by several viruses have been found to colocalize with actin microfilaments, including the TGBp2 movement protein from Potato virus X (PVX), TGBp2 and TGBp3 from Potato mop-top virus, the Hsp70 homolog from Beet yellows virus, as well as both the movement (MP) and 126-kD proteins from Tobacco mosaic virus (TMV; McLean et al., 1995; Haupt et al., 2005; Ju et al., 2005; Liu et al., 2005; Prokhnevsky et al., 2005) In addition, inhibitor studies recently demonstrated that the intracellular trafficking of potato leafroll virus MP to the plasmodesmata (PD) is dependent upon an intact actin cytoskeleton (Vogel et al., 2007). Together, these studies suggest that the trafficking of viral proteins along actin filaments is a mechanism utilized by highly divergent RNA viruses.The only documented example of a plant viral protein found to colocalize with both microfilaments and microtubules in cells is the TMV MP (McLean et al., 1995; reviewed in Beachy and Heinlein, 2000; Lucas, 2006), which has been shown to associate with and stabilize microtubules and contains a motif thought to mimic the region of tubulin responsible for lateral junctions between microtubules (Boyko et al., 2000; Ashby et al., 2006). Interestingly, the CaMV gene II product (P2), an aphid transmission factor, was previously shown by immunoelectron microscopy to associate with microtubules in both insect and plant cells, although the significance of this interaction remains unclear (Blanc et al., 1996). In addition to these two viral proteins found to colocalize with microtubules in planta, the Hsp70 homolog from Beet yellows virus and the coat protein from PVX have both been shown to interact with microtubules in vitro (Karasev et al., 1992; Serazev et al., 2003). Evidence that the intracellular localization of grapevine fanleaf virus MP is disturbed by oryzalin, as well as the finding that the geminivirus replication protein AL1 interacts with a kinesin by yeast two-hybrid assay, may also indicate a potential association of these proteins with microtubules (Kong and Hanley-Bowdoin, 2002; Laporte et al., 2003).In this study, we utilize a fusion between the C terminus of P6 and GFP to visualize P6 inclusions in live cells. We demonstrate that the fusion of P6 with GFP does not interfere with its ability to act as a TAV. We further demonstrate that P6-GFP inclusion bodies move intracellularly and are associated with microtubules, actin microfilaments, and the endoplasmic reticulum (ER). Although P6-GFP inclusion bodies associated with microtubules appear stationary, we show that P6-GFP bodies can traffic along microfilaments and that this movement is severely reduced by treatment with the actin inhibitor latrunculin B (LatB). LatB treatment of N. edwardsonii leaves inhibits the formation of local lesions by CaMV, indicating the potential that P6 trafficking on microfilaments is necessary for CaMV cell-to-cell movement. Additionally, the association of P6-GFP inclusion bodies with microtubules prevents the disruption of microtubules by oryzalin, denoting a tight association between these two proteins. We discuss the potential role of P6 movement and cytoskeletal association in CaMV infection.  相似文献   

20.
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