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.
Therapeutic agents are urgently required to cure several common and fatal neurodegenerative disorders caused by protein misfolding and aggregation, including amyotrophic lateral sclerosis (ALS), Parkinson’s disease (PD), and Alzheimer’s disease (AD). Protein disaggregases that reverse protein misfolding and restore proteins to native structure, function, and localization could mitigate neurodegeneration by simultaneously reversing 1) any toxic gain of function of the misfolded form and 2) any loss of function due to misfolding. Potentiated variants of Hsp104, a hexameric AAA+ ATPase and protein disaggregase from yeast, have been engineered to robustly disaggregate misfolded proteins connected with ALS (e.g., TDP-43 and FUS) and PD (e.g., α-synuclein). However, Hsp104 has no metazoan homologue. Metazoa possess protein disaggregase systems distinct from Hsp104, including Hsp110, Hsp70, and Hsp40, as well as HtrA1, which might be harnessed to reverse deleterious protein misfolding. Nevertheless, vicissitudes of aging, environment, or genetics conspire to negate these disaggregase systems in neurodegenerative disease. Thus, engineering potentiated human protein disaggregases or isolating small-molecule enhancers of their activity could yield transformative therapeutics for ALS, PD, and AD.We urgently need to pioneer game-changing solutions to remedy a number of increasingly prevalent and fatal neurodegenerative disorders, including amyotrophic lateral sclerosis (ALS), Parkinson’s disease (PD), and Alzheimer’s disease (AD; Cushman et al., 2010 ; Jackrel and Shorter, 2015 ). These disorders relentlessly erode our morale and economic resources. Aging is the major risk factor for all of these diseases, which threaten public health on a global scale and represent a severe impediment to living longer lives. A number of promising drugs have emerged to treat cancer and heart disease, but, distressingly, this is not the case for these and other neurodegenerative diseases, for which drug pipelines lie dormant and empty. This situation is unacceptable, and an impending healthcare crisis looms worldwide as population demographics inexorably shift toward older age groups.ALS, PD, AD, and related neurodegenerative disorders are unified by a common underlying theme: the misfolding and aggregation of specific proteins (characteristic for each disease) in the CNS (Cushman et al., 2010 ; Eisele et al., 2015 ). Thus, in ALS, typically an RNA-binding protein with a prion-like domain, TDP-43, mislocalizes from the nucleus to cytoplasmic inclusions in degenerating motor neurons (Neumann et al., 2006 ; Gitler and Shorter, 2011 ; King et al., 2012 ; Robberecht and Philips, 2013 ; March et al., 2016 ). In PD, α-synuclein forms toxic soluble oligomers and cytoplasmic aggregates, termed Lewy bodies, in degenerating dopaminergic neurons (Dehay et al., 2015 ). By contrast, in AD, amyloid-β (Aβ) peptides form extracellular plaques and the microtubule-binding protein, tau, forms cytoplasmic neurofibrillary tangles in afflicted brain regions (Jucker and Walker, 2011 ). Typically, these disorders are categorized into ∼80–90% sporadic cases and ∼10–20% familial cases. Familial forms of disease often have clear genetic causes, which might one day be amenable to gene editing via clustered regularly interspaced short palindromic repeats (CRISPR)-Cas9 therapeutics if critical safety and ethical concerns can be successfully addressed and respected (Doudna and Charpentier, 2014 ; Baltimore et al., 2015 ; Rahdar et al., 2015 ; Callaway, 2016 ). However, the more common sporadic forms of disease often have no clear underlying genetics, and wild-type proteins misfold (Cushman et al., 2010 ; Jucker and Walker, 2011 ; Robberecht and Philips, 2013 ; Dehay et al., 2015 ). Consequently, therapeutic agents that directly target and safely reverse deleterious protein misfolding are likely to have broad utility (Eisele et al., 2015 ).There are no treatments that directly target the reversal of the protein-misfolding phenomena that underlie these disorders (Jackrel and Shorter, 2015 ). Strategies that directly reverse protein misfolding and restore proteins to native form and function could, in principle, eradicate any severely damaging loss-of-function or toxic gain-of-function phenotypes caused by misfolded conformers (Figure 1; Jackrel and Shorter, 2015 ). Moreover, therapeutic disaggregases would dismantle self-templating amyloid or prion structures, which spread pathology and nucleate formation of neurotoxic, soluble oligomers (Figure 1; Cushman et al., 2010 ; Cohen et al., 2013 ; Guo and Lee, 2014 ; Jackrel and Shorter, 2015 ). My group has endeavored to engineer and evolve Hsp104, a hexameric AAA+ ATPase and protein disaggregase from yeast (DeSantis and Shorter, 2012 ; Sweeny and Shorter, 2015 ), to more effectively disaggregate misfolded proteins connected with various neurodegenerative disorders, including ALS (e.g., TDP-43 and FUS) and PD (e.g., α-synuclein). Although wild-type Hsp104 can disaggregate diverse amyloid and prion conformers, as well as toxic soluble oligomers (Lo Bianco et al., 2008 ; DeSantis et al., 2012 ), its activity against human neurodegenerative disease proteins is suboptimal. Is it even possible to improve on existing Hsp104 disaggregase activity, which has been wrought over the course of millions of years of evolution?Open in a separate windowFIGURE 1:Therapeutic protein disaggregases. Two malicious problems are commonly associated with protein misfolding into disordered aggregates, toxic oligomers, and cross–β amyloid or prion fibrils: 1) a toxic gain of function of the protein in various misfolded states; and 2) a loss of function of the protein in the various misfolded states. These problems can contribute to the etiology of diverse neurodegenerative diseases in a combinatorial or mutually exclusive manner. A therapeutic protein disaggregase would reverse protein misfolding and recover natively folded functional proteins from disordered aggregates, toxic oligomers, and cross–β amyloid or prion fibrils. In this way, any toxic gain of function or toxic loss of function caused by protein misfolding would be simultaneously reversed. Ideally, all toxic misfolded conformers would be purged. Therapeutic protein disaggregases could thus have broad utility for various fatal neurodegenerative diseases.Remarkably, the answer to this question is yes! We used nimble yeast models of neurodegenerative proteinopathies (Outeiro and Lindquist, 2003 ; Gitler, 2008 ; Johnson et al., 2008 ; Sun et al., 2011 ; Khurana et al., 2015 ) as a platform to isolate enhanced disaggregases from large libraries of Hsp104 variants generated by error-prone PCR (Jackrel et al., 2014b ). In this way, we reprogrammed Hsp104 to yield the first disaggregases that reverse TDP-43, FUS (another RNA-binding protein with a prion-like domain connected to ALS), and α-synuclein (connected to PD) aggregation and proteotoxicity (Jackrel et al., 2014a , 2015 ; Jackrel and Shorter, 2015 ; Torrente et al., 2016 ). Remarkably, a therapeutic gain of Hsp104 function could be elicited by a single missense mutation (Jackrel et al., 2014a , 2015 ; Jackrel and Shorter, 2015 ). Under conditions in which Hsp104 was ineffective, potentiated Hsp104 variants dissolved protein inclusions, restored protein localization (e.g., TDP-43 returned to the nucleus from cytoplasmic inclusions), suppressed proteotoxicity, and attenuated dopaminergic neurodegeneration in a Caenorhabditis elegans PD model (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ). Remarkably, these therapeutic modalities originated from degenerate loss of amino acid identity at select positions of Hsp104 rather than specific mutation (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ). Some of these changes were remarkably small (Jackrel et al., 2014a , 2015 ; Jackrel and Shorter, 2015 ). Thus, potentiated Hsp104 variants could be generated by removal of a methyl group from a single side chain or addition or removal of a single methylene bridge from a single side chain (Jackrel et al., 2014a , 2015 ; Jackrel and Shorter, 2015 ). Thus, small molecules that bind in accessible regions of Hsp104 rich in potentiating mutations might also be able to enhance activity. However, a small-scale screen for small-molecule modulators of Hsp104 revealed only inhibitors (Torrente et al., 2014 ). Nonetheless, our work has established that disease-associated aggregates and amyloid are tractable targets and that enhanced artificial disaggregases can restore proteostasis and mitigate neurodegeneration (Jackrel and Shorter, 2015 ).One surprising aspect of this work is just how many Hsp104 variants we could isolate with potentiated activity. We now have hundreds (Jackrel et al., 2014a ; Jackrel et al., 2015 ). Typically, potentiated Hsp104 variants displayed enhanced activity against several neurodegenerative disease proteins. For example, Hsp104A503S rescued the aggregation and toxicity of TDP-43, FUS, TAF15, and α-synuclein (Jackrel et al., 2014a ; Jackrel and Shorter, 2014 ). By contrast, some potentiated Hsp104 variants rescued only the aggregation and toxicity of a subset of disease proteins. For example, Hsp104D498V rescued only the aggregation and toxicity of FUS and α-synuclein (Jackrel et al., 2014a ). A challenge that lies ahead is to engineer potentiated Hsp104 variants that are highly substrate specific to mitigate any potential off-target effects, should they arise (Jackrel and Shorter, 2015 ).Very small changes in primary sequence yield potentiated Hsp104 variants. However, Hsp104 has no metazoan homologue (Erives and Fassler, 2015 ). Now comes the important point. Neuroprotection could be broadly achieved by making very subtle modifications to existing human chaperones with newly appreciated disaggregase activity—for example, Hsp110, Hsp70, and Hsp40 (Torrente and Shorter, 2013 ) and HtrA1 (Poepsel et al., 2015 ).Whether Metazoa even possess a powerful protein disaggregation and reactivation machinery had been a long-standing enigma (Torrente and Shorter, 2013 ). However, it has recently emerged that two metazoan chaperone systems—1) Hsp110, Hsp70, and Hsp40 (Torrente and Shorter, 2013 ) and 2) HtrA1 (Poepsel et al., 2015 )—possess disaggregase activity that could be therapeutically harnessed or stimulated to reverse deleterious protein misfolding in neurodegenerative disease. I suspect that Metazoa harbor additional disaggregase systems that remain to be identified (Guo et al., 2014 ). Whether due to vicissitudes of aging, environment, or genetic background, these disaggregase systems fail in the context of ALS, PD, and AD. Based on the surprising precedent of our potentiated versions of Hsp104 (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ), I hypothesize that it is possible to engineer and evolve potentiated variants of these human protein disaggregases to more effectively counter deleterious misfolding events in ALS, PD, and AD (Torrente and Shorter, 2013 ; Mack and Shorter, 2016 ).Using classical biochemical reconstitution, it was discovered that one mammalian protein-disaggregase system comprises three molecular chaperones—Hsp110, Hsp70, and Hsp40—which synergize to dissolve and reactivate model proteins trapped in disordered aggregates and can even depolymerize amyloid fibrils formed by α-synuclein from their ends (Shorter, 2011 ; Duennwald et al., 2012 ; Torrente and Shorter, 2013 ). Hsp110, Hsp70, and Hsp40 isoforms are found in the cytoplasm, nucleus, and endoplasmic reticulum, which suggest that protein disaggregation and reactivation can occur in several compartments (Finka et al., 2015 ). Subsequent studies suggest that this system may be more powerful than initially anticipated (Rampelt et al., 2012 ; Mattoo et al., 2013 ; Gao et al., 2015 ; Nillegoda et al., 2015 ). Nonetheless, this system must become overwhelmed in neurodegenerative disorders. Perhaps selectively vulnerable neurons display particular deficits in this machinery. Hence, potentiating the activity of this system via engineering could be extremely valuable. It is promising that directed evolution studies yielded DnaK (Hsp70 in Escherichia coli) variants with improved ability to refold specific substrates (Aponte et al., 2010 ; Schweizer et al., 2011 ; Mack and Shorter, 2016 ), but whether this can be extended to human Hsp70 and the disaggregation of neurodegenerative disease proteins is uncertain.It is exciting that recent studies have revealed that HtrA1, a homo-oligomeric PDZ serine protease, can dissolve and degrade AD-linked tau and Aβ42 fibrils in an ATP-independent manner (Tennstaedt et al., 2012 ; Poepsel et al., 2015 ). HtrA1 first dissolves tau and Aβ42 fibrils and then degrades them, as protease-defective HtrA1 variants dissolve fibrils without degrading them (Poepsel et al., 2015 ). HtrA1 is found in the cytoplasm (∼30%) but is also secreted (∼70%; Poepsel et al., 2015 ). Indeed, HtrA1 is known to degrade substrates in both the extracellular space and the cytoplasm (Chien et al., 2009 ; Campioni et al., 2010 ; Tiaden and Richards, 2013 ). Thus HtrA1 could dissolve Aβ42 fibrils in the extracellular space and tau fibrils in the cytoplasm and simultaneously destroy the two cardinal features of AD (Poepsel et al., 2015 ). I suspect that this system becomes overwhelmed or is insufficient in AD, and thus potentiating and tailoring HtrA1 disaggregase activity could be a valuable therapeutic strategy. For example, it might be advantageous to simply degrade Aβ42 after disaggregation if the peptide has no beneficial function. Thus HtrA1 variants with enhanced disaggregation and degradation activity against Aβ42 could be extremely useful. However, Aβ42 (and related Aβ peptides) may have physiological functions that are presently underappreciated (Soscia et al., 2010 ; Fedele et al., 2015 ), in which case HtrA1 variants with enhanced disaggregase activity but reduced proteolytic activity could be vital. HtrA1 variants with enhanced disaggregase activity but reduced proteolytic activity may also be particularly important to recover functional tau from neurofibrillary tangles to reverse loss of tau function in AD and various tauopathies (Santacruz et al., 2005 ; Trojanowski and Lee, 2005 ; Dixit et al., 2008 ).I suggest that relatively small changes in primary sequence will yield large increases in disaggregase activity for these systems as they do for Hsp104 (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ). If true, this would further suggest that small molecules that bind in the appropriate regions of Hsp110, Hsp70, Hsp40, or HtrA1 might also enhance disaggregase activity. Thus, isolating small-molecule enhancers of the Hsp110, Hsp70, and Hsp40 or HtrA1 disaggregase systems could yield important therapeutics. Indeed, I hypothesize that enhancing the activity of the Hsp110, Hsp70, and Hsp40 or HtrA1 disaggregase system with specific small molecules will enable dissolution of toxic oligomeric and amyloid forms of various disease proteins and confer therapeutic benefits in ALS, PD, AD, and potentially other neurodegenerative disorders.It is intriguing that several small molecules are already known to enhance various aspects of Hsp70 chaperone activity (Pratt et al., 2015 ; Shrestha et al., 2016 ). These include MKT-077, JG-98, YM-1, YM-8, and 115-7c (Wisen et al., 2010 ; Pratt et al., 2015 ). It is not known whether any of these stimulates the disaggregase activity of the Hsp110, Hsp70, and Hsp40 system. MKT-077, JG-98, YM-1, and YM-8 are rhodocyanines that bind with low micromolar affinity to the nucleotide-binding domain of ADP- but not ATP-bound Hsp70, stabilizing the ADP-bound state (Pratt et al., 2015 ). The ADP-bound state of Hsp70 engages clients with higher affinity, and consequently MKT-077, JG-98, and YM-1 activate binding of Hsp70 to misfolded proteins (Wang et al., 2013 ; Pratt et al., 2015 ). Thus, under some conditions, these small molecules can promote folding of certain Hsp70 clients (Morishima et al., 2011 ; Pratt et al., 2015 ). However, prolonged interaction of clients with Hsp70 promotes their CHIP-dependent ubiquitylation and degradation in vivo (Morishima et al., 2011 ; Wang et al., 2013 ; Pratt et al., 2015 ). Intriguingly, YM-1 promotes clearance of polyglutamine oligomers and aggregates in cells (Wang et al., 2013 ; Pratt et al., 2015 ). MKT-0777, YM-1, JG-98, and YM-8 also promote clearance of tau and confer therapeutic benefit in tauopathy models (Abisambra et al., 2013 ; Miyata et al., 2013 ; Fontaine et al., 2015 ). Of importance, YM-8 is long lived in vivo and crosses the blood–brain barrier (Miyata et al., 2013 ). The dihydropyrimidine 115-7c activates Hsp70 ATPase turnover rate, promotes Hsp70 substrate refolding, and reduces α-synuclein aggregation in cell culture (Wisen et al., 2010 ; Kilpatrick et al., 2013 ). It binds to the IIA subdomain of Hsp70 and promotes the active Hsp70–Hsp40 complex (Wisen et al., 2010 ). Small-molecule enhancers of HtrA1 protease activity have also emerged (Jo et al., 2014 ). Thus it will be important to assess whether these small molecules enhance the activity of their respective disaggregases against various neurodegenerative substrates.Although small molecules that enhance disaggregase activity of endogenous human proteins might be the most immediately translatable, gene-, mRNA-, or protein-based therapies can also be envisioned. For example, adeno-associated viruses expressing enhanced disaggregases might be used to target degenerating neurons (Dong et al., 2005 ; Lo Bianco et al., 2008 ; Deverman et al., 2016 ). Alternatively, if viral vectors are undesirable, modified mRNAs might serve as an alternative to DNA-based gene therapy (Kormann et al., 2011 ). Protein-based therapeutics could also be explored. For example, intraperitoneal injection of human Hsp70 increased lifespan, delayed symptom onset, preserved motor function, and prolonged motor neuron viability in a mouse model of ALS (Gifondorwa et al., 2007 ; Gifondorwa et al., 2012 ). Several other studies suggest that exogenous delivery of Hsp70 can have beneficial, neuroprotective effects in mice (Nagel et al., 2008 ; Bobkova et al., 2014 ; Bobkova et al., 2015 ).Ultimately, if safety and ethical concerns can be overcome in a circumspect, risk-averse manner, CRISPR-Cas9–based therapeutics might even be used to genetically alter the underlying disaggregase to a potentiated form in selectively vulnerable neuronal populations. This approach might be particularly valuable if enhanced disaggregase activity is not detrimental in the long term. Moreover, stem cell–based therapies for replacing lost neurons could also be fortified to express enhanced disaggregase systems. Thus they would be endowed with resistance to potential infection by prion-like conformers that might have accumulated during disease progression (Cushman et al., 2010 ).Enhanced disaggregase activity is likely to be highly advantageous to neurons under circumstances in which protein misfolding has overwhelmed the system (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ). However, inappropriate hyperactivity of protein disaggregases might also have detrimental, off-target effects under regular conditions in which protein misfolding is not an overwhelming issue (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ). Thus it may be advantageous to engineer enhanced protein disaggregases to be highly substrate specific. In this way, off-target effects would be readily avoided. There is strong precedent for directed evolution or engineering of specialized chaperone or protein activity from a generalist antecedent (Wang et al., 2002 ; Farrell et al., 2007 ; Smith et al., 2015 ). Thus, engineering specialist disaggregases for each disease substrate could be achieved. Alternatively, transient or intermittent doses of enhanced disaggregases at specific times or places where they are most needed would also minimize potentially toxic side effects. For example, enhanced disaggregase activity might be applied ephemerally to clear existing misfolded conformers and then be withdrawn once the endogenous proteostasis network regains control. Similarly, it is straightforward to envision administration of small-molecule enhancers of disaggregase activity in intermittent protocols that enable facile recovery from potential side effects (Fontaine et al., 2015 ). In this way, any adverse effects of enhanced protein-disaggregase activity under normal physiological conditions would be avoided. Many barriers will need to be safely overcome to implement a successful therapeutic disaggregase, including how to deliver enhanced disaggregase activity to exactly where it is needed. However, these obstacles are not a reason to be pessimistic. On the contrary, the isolation of engineered disaggregases that efficaciously reverse deleterious misfolding of neurodegenerative disease proteins directs our attention to considerably expand the environs in which they should be sought. My closing sentences, therefore, are intended to be provocative.I suspect that neuroprotection could be broadly actualized via precise but subtle alterations to existing protein-disaggregase modalities. The engineering and evolution of protein disaggregases could yield important solutions to avert an imminent plague of neurodegenerative disorders that promises to devastate our society. I strongly suspect that cures for various neurodegenerative disorders will be realized by pioneering as-yet-uncharted regions of disaggregase sequence space or chemical space to elucidate small-molecule enhancers of disaggregase activity.  相似文献   

2.
Insulin-dependent translocation of glucose transporter 4 (Glut4) to the plasma membrane of fat and skeletal muscle cells plays the key role in postprandial clearance of blood glucose. Glut4 represents the major cell-specific component of the insulin-responsive vesicles (IRVs). It is not clear, however, whether the presence of Glut4 in the IRVs is essential for their ability to respond to insulin stimulation. We prepared two lines of 3T3-L1 cells with low and high expression of myc7-Glut4 and studied its translocation to the plasma membrane upon insulin stimulation, using fluorescence-assisted cell sorting and cell surface biotinylation. In undifferentiated 3T3-L1 preadipocytes, translocation of myc7-Glut4 was low regardless of its expression levels. Coexpression of sortilin increased targeting of myc7-Glut4 to the IRVs, and its insulin responsiveness rose to the maximal levels observed in fully differentiated adipocytes. Sortilin ectopically expressed in undifferentiated cells was translocated to the plasma membrane regardless of the presence or absence of myc7-Glut4. AS160/TBC1D4 is expressed at low levels in preadipocytes but is induced in differentiation and provides an additional mechanism for the intracellular retention and insulin-stimulated release of Glut4.Adipocytes, skeletal muscle cells, and some neurons respond to insulin stimulation by translocating intracellular glucose transporter 4 (Glut4) to the plasma membrane. In all these cells, the insulin-responsive pool of Glut4 is localized in small membrane vesicles, the insulin-responsive vesicles (IRVs; Kandror and Pilch, 2011 ; Bogan, 2012 ). The protein composition of these vesicles has been largely characterized (Kandror and Pilch, 2011 ; Bogan, 2012 ). The IRVs consist predominantly of Glut4, insulin-responsive aminopeptidase (IRAP), sortilin, low-density-lipoprotein receptor–related protein 1 (LRP1), SCAMPs, and VAMP2. Glut4, IRAP, and sortilin physically interact with each other, which might be important for the biogenesis of the IRVs (Shi and Kandror, 2007 ; Shi et al., 2008 ). In addition, the IRVs compartmentalize recycling receptors, such as the transferrin receptor and the IGF2/mannose 6-phosphate receptor, although it is not clear whether these receptors represent obligatory vesicular components or their presence in the IRVs is explained by mass action (Pilch, 2008 ), inefficient sorting, or other reasons.Deciphering of the protein composition of the IRVs is important because it is likely to explain their unique functional property: translocation to the plasma membrane in response to insulin stimulation. Even if we presume that IRV trafficking is controlled by loosely associated peripheral membrane proteins, the latter should still somehow recognize the core vesicular components that create the “biochemical individuality” of this compartment. In spite of our knowledge of the IRV protein composition, however, the identity of the protein(s) that confer insulin sensitivity to these vesicles is unknown.Insulin responsiveness of the IRVs was associated with either IRAP or Glut4. Thus it was shown that Glut4 interacted with the intracellular anchor TUG (Bogan et al., 2003 , 2012 ), whereas IRAP associated with other proteins implemented in the regulation of Glut4 translocation, such as AS160 (Larance et al., 2005 ; Peck et al., 2006 ), p115 (Hosaka et al., 2005 ), tankyrase (Yeh et al., 2007 ), and several others (reviewed in Bogan, 2012 ). Results of these studies, or at least their interpretations, are not necessarily consistent with each other, as the existence of multiple independent anchors for the IRVs is, although possible, unlikely.Ablation of the individual IRV proteins has also led to controversial data. Thus knockout of IRAP decreases total protein levels of Glut4 but does not affect its translocation in the mouse model (Keller et al., 2002 ). On the contrary, knockdown of IRAP in 3T3-L1 adipocytes has a strong inhibitory effect on translocation of Glut4 (Yeh et al., 2007 ). In yet another study, knockdown of IRAP in 3T3-L1 adipocytes did not affect insulin-stimulated translocation of Glut4 but increased its plasma membrane content under basal conditions (Jordens et al., 2010 ). By the same token, total or partial ablation of Glut4 had various effects on expression levels, intracellular localization, and translocation of IRAP (Jiang et al., 2001 ; Abel et al., 2004 ; Carvalho et al., 2004 ; Gross et al., 2004 ; Yeh et al., 2007 ). Knockdown of either sortilin or LRP1 decreased protein levels of Glut4 (Shi and Kandror, 2005 ; Jedrychowski et al., 2010 ).One model that might explain these complicated and somewhat inconsistent results is that depletion of either major integral protein of the IRVs disrupts the network of interactions between vesicular proteins and thus decreases the efficiency of protein sorting into the IRVs (Kandror and Pilch, 2011 ). Correspondingly, the remaining IRV components that cannot be faithfully compartmentalized in the vesicles are either degraded (Jiang et al., 2001 ; Keller et al., 2002 ; Abel et al., 2004 ; Carvalho et al., 2004 ; Shi and Kandror, 2005 ; Yeh et al., 2007 ; Jedrychowski et al., 2010 ) or mistargeted (Jiang et al., 2001 ; Jordens et al., 2010 ), depending on experimental conditions and types of cells used in these studies. In other words, knockdown of any major IRV component may decrease vesicle formation along with insulin responsiveness. Thus, in spite of a large body of literature, the identity of protein(s) that confer insulin responsiveness to the IRVs is unknown.Here we used a gain-of-function approach to address this question. Specifically, we attempted to “build” functional IRVs in undifferentiated 3T3-L1 preadipocytes by forced expression of the relevant proteins. Undifferentiated preadipocytes do not express Glut4 or sortilin and lack IRVs (ElJack et al., 1999 ; Shi and Kandror, 2005 ; Shi et al., 2008 ). Correspondingly, IRAP, which is expressed in these cells, shows low insulin response (Ross et al., 1998 ; Shi et al., 2008 ). We found that ectopic expression of increasing amounts of Glut4 in undifferentiated preadipocytes does not lead to its marked translocation to the plasma membrane upon insulin stimulation. On the contrary, sortilin expressed in undifferentiated preadipocytes was localized in the IRVs and was translocated to the plasma membrane in response to insulin stimulation. Moreover, upon coexpression with Glut4, sortilin dramatically increased its insulin responsiveness to the levels observed in fully differentiated adipocytes. Thus sortilin may represent the key component of the IRVs, which is responsible not only for the formation of vesicles (Shi and Kandror, 2005 ; Ariga et al., 2008 ; Hatakeyama and Kanzaki, 2011 ), but also for their insulin responsiveness. It is worth noting that sortilin levels are significantly decreased in obese and diabetic humans and mice (Kaddai et al., 2009 ). We thus suggest that sortilin may be a novel and important target in the fight against insulin resistance and diabetes.Our experiments also demonstrate that undifferentiated preadipocytes lack a mechanism for the full intracellular retention of Glut4 that can be achieved by ectopic expression of AS160/TBC1D4.  相似文献   

3.
4.
We report the identification and characterization of ERS-24 (Endoplasmic Reticulum SNARE of 24 kD), a new mammalian v-SNARE implicated in vesicular transport between the ER and the Golgi. ERS24 is incorporated into 20S docking and fusion particles and disassembles from this complex in an ATP-dependent manner. ERS-24 has significant sequence homology to Sec22p, a v-SNARE in Saccharomyces cerevisiae required for transport between the ER and the Golgi. ERS-24 is localized to the ER and to the Golgi, and it is enriched in transport vesicles associated with these organelles.Newly formed transport vesicles have to be selectively targeted to their correct destinations, implying the existence of a set of compartment-specific proteins acting as unique receptor–ligand pairs. Such proteins have now been identified (Söllner et al., 1993a ; Rothman, 1994): one partner efficiently packaged into vesicles, termed a v-SNARE,1 and the other mainly localized to the target compartment, a t-SNARE. Cognate pairs of v- and t-SNAREs, capable of binding each other specifically, have been identified for the ER–Golgi transport step (Lian and Ferro-Novick, 1993; Søgaard et al., 1994), the Golgi–plasma membrane transport step (Aalto et al., 1993; Protopopov et al., 1993; Brennwald et al., 1994) in Saccharomyces cerevisiae, and regulated exocytosis in neuronal synapses (Söllner et al., 1993a ; for reviews see Scheller, 1995; Südhof, 1995). Additional components, like p115, rab proteins, and sec1 proteins, appear to regulate vesicle docking by controlling the assembly of SNARE complexes (Søgaard et al., 1994; Lian et al., 1994; Sapperstein et al., 1996; Hata et al., 1993; Pevsner et al., 1994).In contrast with vesicle docking, which requires compartment-specific components, the fusion of the two lipid bilayers uses a more general machinery derived, at least in part, from the cytosol (Rothman, 1994), which includes an ATPase, the N-ethylmaleimide–sensitive fusion protein (NSF) (Block et al., 1988; Malhotra et al., 1988), and soluble NSF attachment proteins (SNAPs) (Clary et al., 1990; Clary and Rothman, 1990; Whiteheart et al., 1993). Only the assembled v–t-SNARE complex provides high affinity sites for the consecutive binding of three SNAPs (Söllner et al., 1993b ; Hayashi et al., 1995) and NSF. When NSF is inactivated in vivo, v–t-SNARE complexes accumulate, confirming that NSF is needed for fusion after stable docking (Søgaard et al., 1994).The complex of SNAREs, SNAPs, and NSF can be isolated from detergent extracts of cellular membranes in the presence of ATPγS, or in the presence of ATP but in the absence of Mg2+, and sediments at ∼20 Svedberg (20S particle) (Wilson et al., 1992). In the presence of MgATP, the ATPase of NSF disassembles the v–t-SNARE complex and also releases SNAPs. It seems likely that this step somehow initiates fusion.To better understand vesicle flow patterns within cells, it is clearly of interest to identify new SNARE proteins. Presently, the most complete inventory is in yeast, but immunolocalization is difficult in yeast compared with animal cells, and many steps in protein transport have been reconstituted in animal extracts (Rothman, 1992) that have not yet been developed in yeast. Therefore, it is important to create an inventory of SNARE proteins in animal cells. The most unambiguous and direct method for isolating new SNAREs is to exploit their ability to assemble together with SNAPs and NSF into 20S particles and to disassemble into subunits when NSF hydrolyzes ATP. Similar approaches have already been successfully used to isolate new SNAREs implicated in ER to Golgi (Søgaard et al., 1994) and intra-Golgi transport (Nagahama et al., 1996), in addition to the original discovery of SNAREs in the context of neurotransmission (Söllner et al., 1993a ).Using this method, we now report the isolation and detailed characterization of ERS-24 (Endoplasmic Reticulum SNARE of 24 kD), a new mammalian v-SNARE that is localized to the ER and Golgi. ERS-24 is found in transport vesicles associated with the transitional areas of the ER and with the rims of Golgi cisternae, suggesting a role for ERS-24 in vesicular transport between these two compartments.  相似文献   

5.
In 1991, a set of transgenic mouse studies took the fields of cell biology and dermatology by storm in providing the first credible evidence that keratin intermediate filaments play a unique and essential role in the structural and mechanical support in keratinocytes of the epidermis. Moreover, these studies intimated that mutations altering the primary structure and function of keratin filaments underlie genetic diseases typified by cellular fragility. This Retrospective on how these studies came to be is offered as a means to highlight the 25th anniversary of these discoveries.Although intermediate filaments (IFs) have been characterized at some level for a longer period of time (Oshima, 2007 ), they were officially discovered as such as recently as 1968 by Howard Holtzer and colleagues while studying the developing skeletal muscle (Ishikawa et al., 1968 ). The advent of gene cloning methods and monospecific antibody production in the late 1970s and throughout the 1980s led to an explosion of data and knowledge about IFs that established them as a large family of genes and proteins that are individually regulated in a tight and evolutionarily conserved tissue- and differentiation-specific manner. Researchers also uncovered some of the remarkable properties of IFs as purified elements in vitro and in living systems and recognized that they occur in the nucleus as well as in the cytoplasm. In spite of the fast pace of progress during that period, however, it was not possible to produce evidence that spoke unequivocally about the functional importance of IFs in cells and tissues, let alone their role in disease.Beginning in the mid- to late 1980s, pioneering experimentation along two distinct lines was underway in the laboratory of Elaine Fuchs, then at the University of Chicago. The eventual merger of these approaches yielded the first formal insight into IF function in vivo, as well as into their direct involvement in human disease. In an effort to define structure–function relationships with regard to the assembly and network formation properties of IFs, one such approach was the application of systematic deletion mutagenesis to keratin 14 (K14), a type I IF that is expressed with its type II partner keratin 5 (K5) in the progenitor basal layer of the epidermis and related complex epithelia. These studies demonstrated that deleting sequences from either end of the central α-helical rod domain of the K14 protein was deleterious for filament formation in a dominant manner both in transfected cells (Albers and Fuchs, 1987 , 1989 ; Figure 1) and the setting of IF polymerization assays involving purified proteins in vitro (e.g., Coulombe et al., 1990 ). The second key effort in the Fuchs lab in the late 1980s resulted in the demonstration that the proximal 2.5 kb and distal 700 base pairs corresponding respectively to the 5′ upstream and 3′ downstream regions of the cloned human K14 gene were sufficient to confer tissue-specific, that is, K14-like, regulation in transgenic mice in vivo (Vassar et al., 1989 ; Figure 1). This tour de force paved the way for the production of a human K14 gene promoter–based cassette (e.g., Saitou et al., 1995 ) that could reliably direct the expression of any open reading frame in a K14-like manner in transgenic mice. As an aside, this tool has had a profound effect on epithelial and skin biology research.Open in a separate windowFIGURE 1:Schematic representation of the strategy and outcome of the experiments that led to the discovery of keratin function and role in genetic disease. Original figures are reproduced to give a realistic account of the data. (A) Examples of a disrupted keratin filament network in cultured epithelial cells transfected with and expressing a dominantly acting K14 deletion mutant (arrows). (Reproduced from Albers and Fuchs, 1987 , with permission.) (B) Preferential expression of a substance P-epitope–tagged transgenic human K14 protein in the basal layer of tail skin epidermis in mouse, conveying the tissue- and differentiation-specific behavior of the transgene. (Reproduced from Vassar et al., 1989 , with permission.) (C) The two experimental approaches described in A and B were combined to assess the consequences of tissue-specific expression of dominantly acting K14 mutants in skin tissue in vivo. (D) Newborn mouse littermates. The mouse at the top is transgenic (Tg) and expresses a mutated form of K14 in the epidermis. It is showing severe skin blistering (arrows), particularly in its front paws, which are heavily used by mouse newborns to feed from their mother. The bottom mouse is a nontransgenic control showing no such blistering. (E, F) Hematoxylin-eosin–stained skin tissue sections showing the location of subepidermal cleavage within the epidermis of a K14 mutant–expressing transgenic mouse (opposing arrows in E). Cleavage occurs at the level of the basal layer, where the mutant keratin is expressed. Again, this is never seen in control wild-type (Wt) skin (F). Bar, 100 μm (E, F). (D–F are from Coulombe et al., 1991b , with permission.) (G) Leg skin in a patient suffering from the Dowling–Meara form of epidermolysis bullosa simplex. Characteristic of this severe variant of this disease, several skin blisters are often grouped in a herpetiform manner (Fine et al., 1991 ).Subsequent use of the human K14 promoter–based cassette to direct the expression of epitope-tagged and selected deletion mutants of K14 gave rise to transgenic mouse pups that exhibited extensive blistering of the skin preferentially at sites of frictional trauma (Coulombe et al., 1991b ; Vassar et al., 1991 ; Figure 1). Electron microscopy showed that skin blistering occurred secondary to a loss of the integrity of keratinocytes located in the basal layer of the epidermis, that is, the precise site of mutant K14 protein accumulation. Such blistering did not occur in transgenic mice expressing a full-length version of human K14 modified to carry only an epitope tag at the C-terminus at similar or higher levels (Coulombe et al., 1991b ; Vassar et al., 1991 ). In addition, the severity of skin blistering in mutant K14–expressing transgenic pups could be directly related to the extent to which the mutant protein had been shown to disrupt filament assembly in transfected cell assays and in IF reconstitution assays in vitro. For instance, tissue-specific expression of a K14 mutant that could severely disrupt 10-nm filament assembly was associated with whole-body skin blistering and the untimely death of mouse pups and, from a pathology perspective, with “tonofilament clumping” and a paucity of visible keratin IFs in transgenic basal keratinocytes. By comparison, expression of another K14 mutant with a less deleterious effect on 10-nm IF assembly was compatible with the survival of transgenic mouse pups and resulted in skin blistering largely limited to the front paws in newborn mice together with altered organization of keratin IFs in basal keratinocytes of transgenic epidermis in situ, albeit without tonofilament clumping. This initial set of mouse strains thus revealed the existence of a direct link between the so-called “genotype” (i.e., mutant K14 characteristics) and the skin phenotype (Coulombe et al., 1991b ; Vassar et al., 1991 ; Fuchs and Coulombe, 1992 ). Electrophoretic analyses of protein samples confirmed that the K14 mutant proteins acted dominantly to produce such spectacular phenotypes in transgenic mouse skin. Finally, blistering also occurred in the mutant K14–expressing transgenic mice in other stratified epithelia known both to express K14 and experience trauma, notably in the oral mucosa (Coulombe et al., 1991b ; Vassar et al., 1991 ).It is worth celebrating the 25th anniversary of these pioneering experiments for the following two reasons. First, the study of these mice provided the first formal demonstration that keratin IFs play a fundamentally important role in structural support in surface epithelia such as the epidermis and oral mucosa. Without proper IF support, epidermal keratinocytes are rendered fragile and cannot sustain trivial frictional stress (Coulombe et al., 1991b ; Fuchs and Coulombe, 1992 ). The second reason is the observation that the phenotype of these K14 mutant–expressing mice proved eerily similar to those of individuals afflicted with the disease epidermolysis bullosa simplex (EBS), a rare, dominantly inherited and debilitating skin condition in which the epidermis and oral mucosa undergo blistering after exposure to trivial mechanical trauma. As observed in the mouse model, tissue cleavage had been shown to result from the loss of integrity of keratinocytes located in the basal layer (Fine et al., 1991 ). Further, other researchers had previously reported on anomalies in the organization of keratin IFs in the basal epidermal keratinocytes of EBS patients (Anton-Lamprecht, 1983 ; Ito et al., 1991 ) or in cultures of epidermal keratinocytes established from EBS patients (Kitajima et al., 1989 ). The Fuchs laboratory thus teamed up with Amy Paller, a physician-scientist and pediatric dermatologist with deep expertise in genodermatoses, and mutations were soon discovered in the K14 gene of two independent and sporadic cases of a severe variant of the disease known as Dowling–Meara EBS (Coulombe et al., 1991a ; Figure 1). The two mutations were heterozygous missense alleles that affected the very same codon in K14 (Arg-125) and were correctly predicted at the time to correspond to a mutational hot spot in type I keratin genes. The mutations were shown to dominantly disrupt 10-nm IF assembly in vitro and/or in transfected keratinocytes in culture (Coulombe et al., 1991a ). Soon thereafter, a team led by Ervin Epstein at University of California, San Francisco (San Francisco, CA), reported on the use of classical linkage analysis to uncover a missense mutation in the K14 gene of a small pedigree with Koebner-type EBS, a less severe variant of the disease (Bonifas et al., 1991 ). The next year, Birgit Lane and colleagues (Lane et al., 1992 ) reported on the occurrence of mutations in keratin 5 (K5), the formal type II keratin assembly partner for K14 in vivo, in another instance of Dowling–Meara EBS.In the years since 1991, a role in structural support has been formally demonstrated for all classes of IFs (Coulombe et al., 2009 ), including the nuclear-localized lamins (e.g., Lammerding et al., 2004 ). Moreover, we now know of several hundred independent instances of mutations in either K5 or K14 in the setting of the EBS disease, with the vast majority of those consisting of dominantly acting missense alleles (Szeverenyi et al., 2008 ; Human Intermediate Filament Database, www.interfil.org, maintained at the Centre for Molecular Medicine and Bioinformatics Institute, Singapore). We also learned that, as anticipated, EBS largely represents a loss-of-function phenotype, since K14-null mice (Lloyd et al., 1995 ), K14-null individuals (Chan et al., 1994 ; Rugg et al., 1994 ), and K5-null mice (Peters et al., 2001 ) all exhibit an EBS-like skin-blistering phenotype (Coulombe et al., 2009 ). Mutations such as Arg125Cys in K14 markedly compromise the remarkable mechanical properties of keratin filaments (Ma et al., 2001 ), as well as the steady-state dynamics of keratin filaments in transfected keratinocytes in culture (Werner et al., 2004 ). Finally, mutations affecting the coding sequence of IF genes have been shown to underlie >100 diseases affecting the human population (Omary et al., 2004 ; Szeverenyi et al., 2008 ; www.interfil.org). Consistent with the exquisite tissue- and cell type–specific regulation of IF genes, these diseases collectively affect a myriad of tissues and organs and are relevant to nearly all branches of medicine. These observations attest to the importance and profound effect that the generation and characterization of mutant K14–expressing transgenic mice has had for cell biology, epithelial physiology, dermatology, and medicine.Many thoughts spring to mind when reminiscing about my involvement with this body of work. First, this effort was prescient of the power of team science and, in particular, of the potential effect of close collaborations involving biologists and physician-scientists. I learned a great deal and benefited immensely from working closely with many colleagues on this project, including Bob Vassar, Kathryn Albers, Linda Degenstein, Liz Hutton, Anthony Letai, Amy Paller, and, last but not least, my postdoctoral mentor and the laboratory head, Elaine Fuchs. Second, there is no substitute for elements such as innovation, hard work, perseverance, boldness, accountability, and great leadership. Elaine had the vision and created the exceptional circumstances necessary to make this set of discoveries possible, and, of equal importance, she was an integral part of the day-to-day progress and maturation of the entire project. Finally, as we all know, there is an intangible element of luck involved in discovery research. In this instance, a strong argument can be made that the studies highlighted here may not have had such a deep and defining effect had the effort been devoted to any IF other than the K5–K14 keratin pairing.What are some of the lingering issues regarding this specific topic that preoccupy us still, 25 years later? Two challenges loom particularly large. First, we have yet to achieve a satisfactory understanding of how mutations in keratin proteins can cause disease. This is due in part to the lack of an atomic-level understanding of the core structure of IFs (which has been a tough nut to crack; Lee et al., 2012 ), along with the reality that, for any relevant IF gene, there is a broad variety of disease-associated (mostly missense) mutations that pepper their primary structure (www.interfil.org). Second, we have yet to achieve success toward the treatment of EBS or any IF-based disorder. Disease characteristics such as low incidence, a dominantly inherited character, genetic heterogeneity (e.g., broad mutational landscape), and, in the case of EBS and related conditions, an intrinsically high rate of cell turnover within the main target tissue significantly add to the challenge of devising safe and effective therapeutic strategies (Coulombe et al., 2009 ). Although efforts are still underway to foster progress on these two challenging issues, the field as a whole has made significant progress in uncovering a plethora of noncanonical functions of keratin IFs (Hobbs et al., 2016 ) in addition to understanding their regulation, dynamics, and many remarkable properties.  相似文献   

6.
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   

7.
8.
SPA2 encodes a yeast protein that is one of the first proteins to localize to sites of polarized growth, such as the shmoo tip and the incipient bud. The dynamics and requirements for Spa2p localization in living cells are examined using Spa2p green fluorescent protein fusions. Spa2p localizes to one edge of unbudded cells and subsequently is observable in the bud tip. Finally, during cytokinesis Spa2p is present as a ring at the mother–daughter bud neck. The bud emergence mutants bem1 and bem2 and mutants defective in the septins do not affect Spa2p localization to the bud tip. Strikingly, a small domain of Spa2p comprised of 150 amino acids is necessary and sufficient for localization to sites of polarized growth. This localization domain and the amino terminus of Spa2p are essential for its function in mating. Searching the yeast genome database revealed a previously uncharacterized protein which we name, Sph1p (Spa2p homolog), with significant homology to the localization domain and amino terminus of Spa2p. This protein also localizes to sites of polarized growth in budding and mating cells. SPH1, which is similar to SPA2, is required for bipolar budding and plays a role in shmoo formation. Overexpression of either Spa2p or Sph1p can block the localization of either protein fused to green fluorescent protein, suggesting that both Spa2p and Sph1p bind to and are localized by the same component. The identification of a 150–amino acid domain necessary and sufficient for localization of Spa2p to sites of polarized growth and the existence of this domain in another yeast protein Sph1p suggest that the early localization of these proteins may be mediated by a receptor that recognizes this small domain.Polarized cell growth and division are essential cellular processes that play a crucial role in the development of eukaryotic organisms. Cell fate can be determined by cell asymmetry during cell division (Horvitz and Herskowitz, 1992; Cohen and Hyman, 1994; Rhyu and Knoblich, 1995). Consequently, the molecules involved in the generation and maintenance of cell asymmetry are important in the process of cell fate determination. Polarized growth can occur in response to external signals such as growth towards a nutrient (Rodriguez-Boulan and Nelson, 1989; Eaton and Simons, 1995) or hormone (Jackson and Hartwell, 1990a , b ; Segall, 1993; Keynes and Cook, 1995) and in response to internal signals as in Caenorhabditis elegans (Goldstein et al., 1993; Kimble, 1994; Priess, 1994) and Drosophila melanogaster (St Johnston and Nusslein-Volhard, 1992; Anderson, 1995) early development. Saccharomyces cerevisiae undergo polarized growth towards an external cue during mating and to an internal cue during budding. Polarization towards a mating partner (shmoo formation) and towards a new bud site requires a number of proteins (Chenevert, 1994; Chant, 1996; Drubin and Nelson, 1996). Many of these proteins are necessary for both processes and are localized to sites of polarized growth, identified by the insertion of new cell wall material (Tkacz and Lampen, 1972; Farkas et al., 1974; Lew and Reed, 1993) to the shmoo tip, bud tip, and mother–daughter bud neck. In yeast, proteins localized to growth sites include cytoskeletal proteins (Adams and Pringle, 1984; Kilmartin and Adams, 1984; Ford, S.K., and J.R. Pringle. 1986. Yeast. 2:S114; Drubin et al., 1988; Snyder, 1989; Snyder et al., 1991; Amatruda and Cooper, 1992; Lew and Reed, 1993; Waddle et al., 1996), neck filament components (septins) (Byers and Goetsch, 1976; Kim et al., 1991; Ford and Pringle, 1991; Haarer and Pringle, 1987; Longtine et al., 1996), motor proteins (Lillie and Brown, 1994), G-proteins (Ziman, 1993; Yamochi et al., 1994; Qadota et al., 1996), and two membrane proteins (Halme et al., 1996; Roemer et al., 1996; Qadota et al., 1996). Septins, actin, and actin-associated proteins localize early in the cell cycle, before a bud or shmoo tip is recognizable. How this group of proteins is localized to and maintained at sites of cell growth remains unclear.Spa2p is one of the first proteins involved in bud formation to localize to the incipient bud site before a bud is recognizable (Snyder, 1989; Snyder et al., 1991; Chant, 1996). Spa2p has been localized to where a new bud will form at approximately the same time as actin patches concentrate at this region (Snyder et al., 1991). An understanding of how Spa2p localizes to incipient bud sites will shed light on the very early stages of cell polarization. Later in the cell cycle, Spa2p is also found at the mother–daughter bud neck in cells undergoing cytokinesis. Spa2p, a nonessential protein, has been shown to be involved in bud site selection (Snyder, 1989; Zahner et al., 1996), shmoo formation (Gehrung and Snyder, 1990), and mating (Gehrung and Snyder, 1990; Chenevert et al., 1994; Yorihuzi and Ohsumi, 1994; Dorer et al., 1995). Genetic studies also suggest that Spa2p has a role in cytokinesis (Flescher et al., 1993), yet little is known about how this protein is localized to sites of polarized growth.We have used Spa2p green fluorescent protein (GFP)1 fusions to investigate the early localization of Spa2p to sites of polarized growth in living cells. Our results demonstrate that a small domain of ∼150 amino acids of this large 1,466-residue protein is sufficient for targeting to sites of polarized growth and is necessary for Spa2p function. Furthermore, we have identified and characterized a novel yeast protein, Sph1p, which has homology to both the Spa2p amino terminus and the Spa2p localization domain. Sph1p localizes to similar regions of polarized growth and sph1 mutants have similar phenotypes as spa2 mutants.  相似文献   

9.
10.
NAD-isocitrate dehydrogenase (NAD-IDH) from the eukaryotic microalga Chlamydomonas reinhardtii was purified to electrophoretic homogeneity by successive chromatography steps on Phenyl-Sepharose, Blue-Sepharose, diethylaminoethyl-Sephacel, and Sephacryl S-300 (all Pharmacia Biotech). The 320-kD enzyme was found to be an octamer composed of 45-kD subunits. The presence of isocitrate plus Mn2+ protected the enzyme against thermal inactivation or inhibition by specific reagents for arginine or lysine. NADH was a competitive inhibitor (Ki, 0.14 mm) and NADPH was a noncompetitive inhibitor (Ki, 0.42 mm) with respect to NAD+. Citrate and adenine nucleotides at concentrations less than 1 mm had no effect on the activity, but 10 mm citrate, ATP, or ADP had an inhibitory effect. In addition, NAD-IDH was inhibited by inorganic monovalent anions, but l-amino acids and intermediates of glycolysis and the tricarboxylic acid cycle had no significant effect. These data support the idea that NAD-IDH from photosynthetic organisms may be a key regulatory enzyme within the tricarboxylic acid cycle.IDH catalyzes the oxidative decarboxylation of isocitrate to produce 2-oxoglutarate. According to the specificity for the electron acceptor, two enzymes with IDH activity are known, NAD-IDH (EC 1.1.1.41) and NADP-IDH (EC 1.1.1.42) (Chen and Gadal, 1990a).In photosynthetic organisms NADP-IDH has been detected in the cytosol, chloroplasts, mitochondria, and peroxisomes. Cytosolic NADP-IDH has been purified from higher plants (Chen et al., 1988) and eukaryotic algae (Martínez-Rivas et al., 1996), and its cDNA has been cloned from alfalfa (Shorrosh and Dixon, 1992), soybean (Udvardi et al., 1993), potato (Fieuw et al., 1995), and tobacco (Gálvez et al., 1996). This 80-kD isoenzyme is a dimer, and it is likely to be involved in the synthesis of NADPH for biosynthetic purposes in the cytosol (Chen et al., 1988), in the synthesis of 2-oxoglutarate for ammonium assimilation (Chen and Gadal, 1990b), and in the cycling, redistribution, and export of amino acids (Fieuw et al., 1995). Chloroplastic NADP-IDH has been studied in higher plants (Gálvez et al., 1994) and eukaryotic algae (Martínez-Rivas and Vega, 1994). It is a 154-kD dimer that has been proposed to be involved in the supply of NADPH for biosynthetic reactions in the chloroplast when photosynthetic NADPH production is low (Gálvez et al., 1994). The mitochondrial NADP-IDH of higher plants may have a physiological role in the production of NADPH, which can be converted to NADH by a transhydrogenase or used to reduce glutathione in the mitochondrial matrix (Rasmusson and Møller, 1990). NADP-IDH activity has also been detected in peroxisomes from spinach leaves (Yamazaki and Tolbert, 1970).NAD-IDH is localized exclusively in the mitochondria in association with the TCA cycle. This enzyme has been purified from several nonphotosynthetic eukaryotes such as fungi (Keys and McAlister-Henn, 1990; Alvarez-Villafañe et al., 1996) and animals (Giorgio et al., 1970), in which it appears to be a 300-kD octamer. Its key regulatory role in the TCA cycle is well documented. The NAD-IDH from yeast is activated by AMP and citrate (Hathaway and Atkinson, 1963), whereas the animal enzyme is activated by ADP and citrate (Cohen and Colman, 1972). In addition, the NAD-IDH cDNAs have been cloned from yeast (Cupp and McAlister-Henn, 1991, 1992) and animals (Nichols et al., 1995; Zeng et al., 1995). In these organisms, the enzyme is composed of two (yeast) or more (animals) different subunits encoded by different genes.To our knowledge, no NAD-IDH from photosynthetic organisms has yet been purified to homogeneity, mainly because of the low stability of the enzyme (Oliver and McIntosh, 1995). However, partial purifications have been reported from pea (Cox and Davies, 1967; Cox, 1969; McIntosh and Oliver, 1992), potato (Laties, 1983), spruce (Cornu et al., 1996), and the eukaryotic microalga Chlamydomonas reinhardtii (Martínez-Rivas and Vega, 1994). Matrix and membrane forms of the enzyme have been detected in potato (Tezuka and Laties, 1983) and pea (McIntosh, 1997). Although it is an allosteric enzyme that exhibits sigmoidal kinetics with respect to isocitrate (Cox and Davies, 1967; McIntosh and Oliver, 1992) and is activated in vitro by ABA (Tezuka et al., 1990), the regulatory importance of NAD-IDH in photosynthetic organisms is still under debate.To elucidate the regulatory significance of NAD-IDH in photosynthetic organisms and its apparent contribution to the 2-oxoglutarate supply for ammonium assimilation, we have purified and characterized the NAD-IDH from C. reinhardtii.  相似文献   

11.
The desmosome is a highly organized plasma membrane domain that couples intermediate filaments to the plasma membrane at regions of cell–cell adhesion. Desmosomes contain two classes of cadherins, desmogleins, and desmocollins, that bind to the cytoplasmic protein plakoglobin. Desmoplakin is a desmosomal component that plays a critical role in linking intermediate filament networks to the desmosomal plaque, and the amino-terminal domain of desmoplakin targets desmoplakin to the desmosome. However, the desmosomal protein(s) that bind the amino-terminal domain of desmoplakin have not been identified. To determine if the desmosomal cadherins and plakoglobin interact with the amino-terminal domain of desmoplakin, these proteins were co-expressed in L-cell fibroblasts, cells that do not normally express desmosomal components. When expressed in L-cells, the desmosomal cadherins and plakoglobin exhibited a diffuse distribution. However, in the presence of an amino-terminal desmoplakin polypeptide (DP-NTP), the desmosomal cadherins and plakoglobin were observed in punctate clusters that also contained DP-NTP. In addition, plakoglobin and DP-NTP were recruited to cell–cell interfaces in L-cells co-expressing a chimeric cadherin with the E-cadherin extracellular domain and the desmoglein-1 cytoplasmic domain, and these cells formed structures that were ultrastructurally similar to the outer plaque of the desmosome. In transient expression experiments in COS cells, the recruitment of DP-NTP to cell borders by the chimera required co-expression of plakoglobin. Plakoglobin and DP-NTP co-immunoprecipitated when extracted from L-cells, and yeast two hybrid analysis indicated that DP-NTP binds directly to plakoglobin but not Dsg1. These results identify a role for desmoplakin in organizing the desmosomal cadherin–plakoglobin complex and provide new insights into the hierarchy of protein interactions that occur in the desmosomal plaque.Desmosomes are highly organized adhesive intercellular junctions that couple intermediate filaments to the cell surface at sites of cell–cell adhesion (Farquhar and Palade, 1963; Staehelin, 1974; Schwarz et al., 1990; Garrod, 1993; Collins and Garrod, 1994; Cowin and Burke, 1996; Kowalczyk and Green, 1996). Desmosomes are prominent in tissues that experience mechanical stress, such as heart and epidermis, and the disruption of desmosomes or the intermediate filament system in these organs has devastating effects on tissue integrity (Steinert and Bale, 1993; Coulombe and Fuchs, 1994; Fuchs, 1994; McLean and Lane, 1995; Stanley, 1995; Bierkamp et al., 1996; Ruiz et al., 1996). Desmosomes are highly insoluble structures that can withstand harsh denaturing conditions (Skerrow and Matoltsy, 1974; Gorbsky and Steinberg, 1981; Jones et al., 1988; Schwarz et al., 1990). This property of desmosomes facilitated early identification of desmosomal components but has impaired subsequent biochemical analysis of the protein complexes that form between desmosomal components. Ultrastructurally, desmosomes contain a core region that includes the plasma membranes of adjacent cells and a cytoplasmic plaque that anchors intermediate filaments to the plasma membrane. The plaque can be further divided into an outer dense plaque subjacent to the plasma membrane and an inner dense plaque through which intermediate filaments appear to loop.Molecular genetic analysis has revealed that the desmosomal glycoproteins, the desmogleins and desmocollins, are members of the cadherin family of cell–cell adhesion molecules (for review see Buxton et al., 1993, 1994; Cowin and Mechanic, 1994; Kowalczyk et al., 1996). The classical cadherins, such as E-cadherin, mediate calcium-dependent, homophilic cell–cell adhesion (Nagafuchi et al., 1987). The mechanism by which the desmosomal cadherins mediate cell–cell adhesion remains elusive (Amagai et al., 1994; Chidgey et al., 1996; Kowalczyk et al., 1996), although heterophilic interactions have recently been detected between desmogleins and desmocollins (Chitaev and Troyanovsky, 1997). Both classes of the desmosomal cadherins associate with the cytoplasmic plaque protein plakoglobin (Kowalczyk et al., 1994; Mathur et al., 1994; Roh and Stanley, 1995b ; Troyanovsky et al., 1994), which is part of a growing family of proteins that share a repeated motif first identified in the Drosophila protein Armadillo (Peifer and Wieschaus, 1990). This multigene family also includes the desmosomal proteins band 6/plakophilin 1, plakophilin 2a and 2b, and p0071, which are now considered to comprise a subclass of the armadillo family of proteins (Hatzfeld et al., 1994; Heid et al., 1994; Schmidt et al., 1994; Hatzfeld and Nachtsheim, 1996; Mertens et al., 1996).The most abundant desmosomal plaque protein is desmoplakin, which is predicted to be a homodimer containing two globular end domains joined by a central α-helical coiled-coil rod domain (O''Keefe et al., 1989; Green et al., 1990; Virata et al., 1992). Previous studies have demonstrated that the carboxyl-terminal domain of desmoplakin interacts with intermediate filaments (Stappenbeck and Green, 1992; Stappenbeck et al., 1993; Kouklis et al., 1994; Meng et al., 1997), and the amino-terminal domain of desmoplakin is required for desmoplakin localization to the desmosomal plaque (Stappenbeck et al., 1993). Direct evidence supporting a role for desmoplakin in intermediate filament attachment to desmosomes was provided recently when expression of an amino-terminal polypeptide of desmoplakin was found to displace endogenous desmoplakin from cell borders and disrupt intermediate filament attachment to the cell surface in A431 epithelial cell lines (Bornslaeger et al., 1996).The classical cadherins, such as E-cadherin, bind directly to both β-catenin and plakoglobin (Aberle et al., 1994; Jou et al., 1995; for review see Cowin and Burke, 1996). β-Catenin is also an armadillo family member (McCrea et al., 1991; Peifer et al., 1992), and both plakoglobin and β-catenin bind directly to α-catenin (Aberle et al., 1994, 1996; Jou et al., 1995; Sacco et al., 1995; Obama and Ozawa, 1997). α-Catenin is a vinculin homologue (Nagafuchi et al., 1991) and associates with both α-actinin and actin (Knudson et al., 1995; Rimm et al., 1995; Nieset et al., 1997). Through interactions with β- and α-catenin, E-cadherin is coupled indirectly to the actin cytoskeleton, and this linkage is required for the adhesive activity of E-cadherin (Ozawa et al., 1990; Shimoyama et al., 1992). In addition, E-cadherin association with plakoglobin appears to be required for assembly of desmosomes (Lewis et al., 1997), underscoring the importance of E-cadherin in the overall program of intercellular junction assembly. However, the hierarchy of molecular interactions that couple the desmosomal cadherins to the intermediate filament cytoskeleton is largely unknown, although the desmocollin cytoplasmic domain appears to play an important role in recruiting components of the desmosomal plaque (Troyanovsky et al., 1993, 1994). Since desmosomal cadherins form complexes with plakoglobin and because the amino-terminal domain of desmoplakin is required for desmoplakin localization at desmosomes, we hypothesized that the amino-terminal domain of desmoplakin interacts with the desmosomal cadherin– plakoglobin complex.In previous studies, we used L-cell fibroblasts to characterize plakoglobin interactions with the cytoplasmic domains of the desmosomal cadherins and found that the desmosomal cadherins regulate plakoglobin metabolic stability (Kowalczyk et al., 1994) but do not mediate homophilic adhesion (Kowalczyk et al., 1996). To test the ability of the desmoplakin amino-terminal domain to interact with the desmosomal cadherin–plakoglobin complex, we established a series of L-cell lines expressing the desmosomal cadherins in the presence or absence of a desmoplakin amino-terminal polypeptide (DP-NTP).1 The results indicate that one important function of the desmoplakin amino-terminal domain is to cluster desmosomal cadherin–plakoglobin complexes. In addition, DP-NTP and plakoglobin were found to form complexes that could be co-immunoprecipitated from L-cell lysates. Using the yeast two hybrid system, DP-NTP was found to bind directly to plakoglobin but not Dsg1. These data suggest that plakoglobin couples the amino-terminal domain of desmoplakin to the desmosomal cadherins and that desmoplakin plays an important role in organizing the desmosomal cadherin–plakoglobin complex into discrete plasma membrane domains.  相似文献   

12.
13.
The Acanthamoeba castellanii myosin-Is were the first unconventional myosins to be discovered, and the myosin-I class has since been found to be one of the more diverse and abundant classes of the myosin superfamily. We used two-dimensional (2D) crystallization on phospholipid monolayers and negative stain electron microscopy to calculate a projection map of a “classical” myosin-I, Acanthamoeba myosin-IB (MIB), at ∼18 Å resolution. Interpretation of the projection map suggests that the MIB molecules sit upright on the membrane. We also used cryoelectron microscopy and helical image analysis to determine the three-dimensional structure of actin filaments decorated with unphosphorylated (inactive) MIB. The catalytic domain is similar to that of other myosins, whereas the large carboxy-terminal tail domain differs greatly from brush border myosin-I (BBM-I), another member of the myosin-I class. These differences may be relevant to the distinct cellular functions of these two types of myosin-I. The catalytic domain of MIB also attaches to F-actin at a significantly different angle, ∼10°, than BBM-I. Finally, there is evidence that the tails of adjacent MIB molecules interact in both the 2D crystal and in the decorated actin filaments.The myosin superfamily consists of at least 12 distinct classes that vary both in the sequence of their conserved myosin catalytic domains as well as in their unique tails (Mooseker and Cheney, 1995; Sellers and Goodson, 1995). For many years the only known myosins were the double-headed, filament-forming myosins found in muscle (conventional myosins or myosins-II). The remaining classes of myosin have been termed “unconventional myosins” to differentiate them from the myosins-II. Probably the most thoroughly studied class of unconventional myosins is the myosin-I class. These small, single-headed myosins bind to membrane lipids through a basic domain in their tail (for review see Pollard et al., 1991; Mooseker and Cheney, 1995). The first unconventional myosin (and first myosin-I) was isolated from Acanthamoeba castellanii (Pollard and Korn, 1973 a,b), and was purified on the basis of its K+, EDTA, and actin-activated MgATPase activities. However, this myosin was unusual in that it had a single heavy chain of ∼140 kD, in contrast to the two ∼200-kD heavy chains of myosin-II (Pollard and Korn, 1973 a).Three isoforms of the classical Acanthamoeba myosins-I are now known: myosins-IA, -IB, and -IC (Maruta and Korn, 1977a ,b; Maruta et al., 1979). Each of the isoforms consists of a conserved myosin catalytic domain, a binding site for one or two light chains, a basic domain, a GPA1(Q) domain (rich in glycine, proline and alanine [glutamine]), and an scr-homology domain-3 (SH3) domain (Pollard et al., 1991; Mooseker and Cheney, 1995). These myosins-I can associate with membranes or with actin filaments through their tail domains. An electrostatic association of myosin-I with anionic phospholipids and with base-stripped membranes has been shown to occur (Adams and Pollard, 1989; Miyata et al., 1989; Hayden et al., 1990), and this interaction has been mapped to the basic domain (Doberstein and Pollard, 1992). Interestingly, these myosins also contain a second, ATP-insensitive actin binding site (Lynch et al., 1986) enabling them to mediate actin–actin movements (Albanesi et al., 1985; Fujisaki et al., 1985). In myosin-IA (Lynch et al., 1986) and myosin-IC (Doberstein and Pollard, 1992), this binding site was localized to the GPA domain. Acanthamoeba myosins-I have maximal steady-state actin-activated ATPase rates of ∼10–20 s−1 (Pollard and Korn, 1973 b; Albanesi et al., 1983), and an unusual triphasic dependence upon actin concentration (Pollard and Korn, 1973 b; Albanesi et al., 1983). This triphasic activation is due to the actin cross-linking ability imparted by the ATP-insensitive actin binding site on the tail (Albanesi et al., 1985). Analysis of the individual steps in the ATPase cycle by transient kinetics revealed that the mechanism of myosin-IA is similar to slow skeletal muscle myosin, whereas myosin-IB (MIB) is similar to fast skeletal muscle myosin (Ostap and Pollard, 1996). The in vitro motility of myosin-I has also been well characterized (Zot et al., 1992). The maximal rate of filament sliding is ∼0.2 μm s−1. Interestingly, this rate is ∼10–50× slower than the rates observed for skeletal muscle myosin, even though the ATPase rates are comparable.MIB consists of a 125-kD heavy chain and a single 27-kD light chain (Maruta et al., 1979; Jung et al., 1987). This isoform is primarily associated with the plasma membrane as well as vacuolar membranes (Baines et al., 1992). MIB appears to be associated with the plasma membrane at sites of phagocytosis and was concentrated at the tips of pseudopodia (Baines et al., 1992). This localization suggests that MIB may be involved in membrane dynamics at the cell surface. MIB is regulated by heavy chain phosphorylation of serine 411 (Brzeska et al., 1989, 1990), which is located at the actin-binding site (Rayment et al., 1993). Similar to the myosin-I isoforms in Acanthamoeba, heavy chain phosphorylation results in >20-fold activation of the actin-activated myosin-I ATPase activity (Albanesi et al., 1983). This activation is not the result of changes in the binding of myosin-I to F-actin (Albanesi et al., 1983; Ostap and Pollard, 1996). The transient kinetic studies of Ostap and Pollard (1996) suggest that phosphorylation regulates the rate-limiting phosphate release step, the transition from weakly bound intermediates in rapid equilibrium with actin to strongly bound states, capable of sustaining force.Despite the extensive analysis of ameboid myosin-I biochemical properties and in vivo function, there is little structural information on these myosins. The only detailed structural information available for the myosins-I comes from recent electron microscopy studies on brush border myosin-I (BBM-I) (Jontes et al., 1995; Jontes and Milligan, 1997a ,b; Whittaker and Milligan, 1997), a structurally distinct myosin-I subtype. Therefore, we investigated the structure of a “classical,” ameboid-type myosin, Acanthamoeba MIB using electron microscopy. First, electron micrographs of negatively stained two-dimensional (2D) crystals were used to generate a projection map of MIB at ∼18 Å resolution. In addition, we used cryoelectron microscopy and helical image analysis to produce a moderate resolution three-dimensional (3D) map (30 Å) of actin filaments decorated with MIB. These studies enabled us to compare the structure of MIB with BBM-I. The comparison of MIB with BBM-I reveals marked structural differences in the tail domains of these two proteins; MIB appears to have a much shorter “lever arm” and a more compact tail, whereas most of the BBM-I mass is composed of an extended light chain–binding domain (LCBD). In addition, the MIB catalytic domain appears to be slightly tilted compared to BBM-I, with respect to the F-actin axis. Our structural results suggest that these two types of myosin-I may have distinct intracellular functions.  相似文献   

14.
This paper presents evidence that a member of the L1 family of ankyrin-binding cell adhesion molecules is a substrate for protein tyrosine kinase(s) and phosphatase(s), identifies the highly conserved FIGQY tyrosine in the cytoplasmic domain as the principal site of phosphorylation, and demonstrates that phosphorylation of the FIGQY tyrosine abolishes ankyrin-binding activity. Neurofascin expressed in neuroblastoma cells is subject to tyrosine phosphorylation after activation of tyrosine kinases by NGF or bFGF or inactivation of tyrosine phosphatases with vanadate or dephostatin. Furthermore, both neurofascin and the related molecule Nr-CAM are tyrosine phosphorylated in a developmentally regulated pattern in rat brain. The FIGQY sequence is present in the cytoplasmic domains of all members of the L1 family of neural cell adhesion molecules. Phosphorylation of the FIGQY tyrosine abolishes ankyrin binding, as determined by coimmunoprecipitation of endogenous ankyrin and in vitro ankyrin-binding assays. Measurements of fluorescence recovery after photobleaching demonstrate that phosphorylation of the FIGQY tyrosine also increases lateral mobility of neurofascin expressed in neuroblastoma cells to the same extent as removal of the cytoplasmic domain. Ankyrin binding, therefore, appears to regulate the dynamic behavior of neurofascin and is the target for regulation by tyrosine phosphorylation in response to external signals. These findings suggest that tyrosine phosphorylation at the FIGQY site represents a highly conserved mechanism, used by the entire class of L1-related cell adhesion molecules, for regulation of ankyrin-dependent connections to the spectrin skeleton.Vertebrate L1, neurofascin, neuroglial cell adhesion molecule (Ng-CAM),1 Ng-CAM–related cell adhesion molecule (Nr-CAM), and Drosophila neuroglian are members of a family of nervous system cell adhesion molecules that possess variable extracellular domains comprised of Ig and fibronectin type III domains and a relatively conserved cytoplasmic domain (Grumet, 1991; Hortsch and Goodman, 1991; Rathgen and Jessel, 1991; Sonderegger and Rathgen, 1992; Hortsch, 1996). Members of this family, including a number of alternatively spliced forms, are abundant in the nervous system during early development as well as in adults. Neurofascin and Nr-CAM, for example, constitute ∼0.5% of the total membrane protein in adult brain (Davis et al., 1993; Davis and Bennett, 1994). Cellular functions attributed to the L1 family include axon fasciculation (Stallcup and Beasley, 1985; Landmesser et al., 1988; Brummendorf and Rathjen, 1993; Bastmeyer et al., 1995; Itoh et al., 1995; Magyar-Lehmann et al., 1995), axonal guidance (van den Pol and Kim, 1993; Liljelund et al., 1994; Brittis and Silver, 1995; Brittis et al., 1995; Lochter et al., 1995; Wong et al., 1996), neurite extension (Chang et al., 1987; Felsenfeld et al., 1994; Hankin and Lagenaur, 1994; Ignelzi et al., 1994; Williams et al., 1994a ,b,c,d; Doherty et al., 1995; Zhao and Siu, 1995), a role in long term potentiation (Luthl et al., 1994), synaptogenesis (Itoh et al., 1995), and myelination (Wood et al., 1990). The potential clinical importance of this group of proteins has been emphasized by the findings that mutations in the L1 gene on the X chromosome are responsible for developmental anomalies including hydrocephalus and mental retardation (Rosenthal et al., 1992; Jouet et al., 1994; Wong et al., 1995).The conserved cytoplasmic domains of L1 family members include a binding site for the membrane skeletal protein ankyrin. This interaction was first described for neurofascin (Davis et. al., 1993) and subsequently has been observed for L1, Nr-CAM (Davis and Bennett, 1994), and Drosophila neuroglian (Dubreuil et al., 1996). The membrane-binding domain of ankyrin contains two distinct sites for neurofascin and has the potential to promote lateral association of neurofascin and presumably other L1 family members (Michaely and Bennett, 1995). Nodes of Ranvier are physiologically relevant axonal sites where ankyrin and L1 family members collaborate, based on findings of colocalization of a specialized isoform of ankyrin with alternatively spliced forms of neurofascin and NrCAM in adults (Davis et al., 1996) as well as in early axonal developmental intermediates (Lambert, S., J. Davis, P. Michael, and V. Bennett. 1995. Mol. Biol. Cell. 6:98a).L1, after homophilic and/or heterophilic binding, participates in signal transduction pathways that ultimately are associated with neurite extension and outgrowth (Ignelzi et al., 1994; Williams et al., 1994a ,b,c,d; Doherty et al., 1995). L1 copurifies with a serine–threonine protein kinase (Sadoul et al., 1989) and is phosphorylated on a serine residue that is not conserved among other family members (Wong et al., 1996). L1 pathway(s) may also involve G proteins, calcium channels, and tyrosine phosphorylation (Williams et al., 1994a ,b,c,d; Doherty et al., 1995). After homophilic interactions, L1 directly activates a tyrosine signaling cascade after a lateral association of its ectodomain with the fibroblast growth factor receptor (Doherty et al., 1995). Antibodies against L1 have also been shown to activate protein tyrosine phosphatase activity in growth cones (Klinz et al., 1995). However, details of the downstream substrates of L1-promoted phosphorylation and dephosphorylation and possible roles of the cytoplasmic domain are not known.Tyrosine phosphorylation is well established to modulate cell–cell and cell–extracellular matrix interactions involving integrins and their associated proteins (Akiyama et al., 1994; Arroyo et al., 1994; Schlaepfer et al., 1994; Law et al., 1996) as well as the cadherins (Balsamo et al., 1996; Krypta et al., 1996; Brady-Kalnay et al., 1995; Shibamoto et al., 1995; Hoschuetzky et al., 1994; Matsuyoshi et al., 1992). For example, the adhesive functions of the calciumdependent cadherin cell adhesion molecule are mediated by a dynamic balance between tyrosine phosphorylation of β-catenin by TrkA and dephosphorylation via the LARtype protein tyrosine phosphatase (Krypta et al., 1996). In this example the regulation of binding among the structural proteins is the result of a coordination between classes of protein kinases and protein phosphatases.This study presents evidence that neurofascin, expressed in a rat neuroblastoma cell line, is a substrate for both tyrosine kinases and protein tyrosine phosphatases at a tyrosine residue conserved among all members of the L1 family. Site-specific tyrosine phosphorylation promoted by both tyrosine kinase activators (NGF and bFGF) and protein tyrosine phosphatase inhibitors (dephostatin and vanadate) is a strong negative regulator of the neurofascin– ankyrin binding interaction and modulates the membrane dynamic behavior of neurofascin. Furthermore, neurofascin and, to a lesser extent Nr-CAM, are also shown here to be tyrosine phosphorylated in developing rat brain, implying a physiological relevance to this phenomenon. These results indicate that neurofascin may be a target for the coordinate control over phosphorylation that is elicited by protein kinases and phosphatases during in vivo tyrosine phosphorylation cascades. The consequent decrease in ankyrin-binding capacity due to phosphorylation of neurofascin could represent a general mechanism among the L1 family members for regulation of membrane–cytoskeletal interactions in both developing and adult nervous systems.  相似文献   

15.
16.
The importin-α/β complex and the GTPase Ran mediate nuclear import of proteins with a classical nuclear localization signal. Although Ran has been implicated also in a variety of other processes, such as cell cycle progression, a direct function of Ran has so far only been demonstrated for importin-mediated nuclear import. We have now identified an entire class of ∼20 potential Ran targets that share a sequence motif related to the Ran-binding site of importin-β. We have confirmed specific RanGTP binding for some of them, namely for two novel factors, RanBP7 and RanBP8, for CAS, Pse1p, and Msn5p, and for the cell cycle regulator Cse1p from Saccharomyces cerevisiae. We have studied RanBP7 in more detail. Similar to importin-β, it prevents the activation of Ran''s GTPase by RanGAP1 and inhibits nucleotide exchange on RanGTP. RanBP7 binds directly to nuclear pore complexes where it competes for binding sites with importin-β, transportin, and apparently also with the mediators of mRNA and U snRNA export. Furthermore, we provide evidence for a Ran-dependent transport cycle of RanBP7 and demonstrate that RanBP7 can cross the nuclear envelope rapidly and in both directions. On the basis of these results, we propose that RanBP7 might represent a nuclear transport factor that carries an as yet unknown cargo, which could apply as well for this entire class of related RanGTP-binding proteins.The nuclear pore complexes (NPC)1 are the sites where the exchange of macromolecules between nucleus and cytoplasm occurs (Feldherr et al., 1984). Transport through the NPCs is bidirectional and comprises a multitude of substrates. Small molecules can passively diffuse through the NPC. The transport of proteins and RNAs >40–60 kD is, however, generally an active process, i.e., it is energy dependent (Newmeyer et al., 1986) and mediated by saturable transport receptors (Goldfarb et al., 1986; Michaud and Goldfarb, 1991; Jarmolowski et al., 1994). To accomplish multiple rounds of transport, these transport receptors are thought to shuttle between nucleus and cytoplasm (Goldfarb et al., 1986). An import receptor, for example, has to bind its import substrate initially in the cytoplasm, release it after NPC passage into the nucleus, and return to the cytoplasm without the cargo. Conversely, an export factor has to bind the export substrate only in the nucleus and on the way out. This model predicts asymmetry in these transport cycles and implies that the binding of the transport receptor to its cargo is regulated by the different environments of nucleus and cytoplasm.The nuclear import of proteins with a classical nuclear localization signal (NLS) is to date the best characterized nucleocytoplasmic transport pathway (for reviews see Görlich and Mattaj, 1996; Koepp and Silver, 1996; Schlenstedt, 1996). The signal contains one or more clusters of basic amino acids (for review see Dingwall and Laskey, 1991) and is recognized by the importin-α/β complex (for references see Sweet and Gerace, 1995; Panté and Aebi, 1996). The α subunit provides the NLS binding site (Imamoto et al., 1995; Weis et al., 1995) and is therefore also called the NLS receptor (Adam and Gerace, 1991). The β subunit accounts for the interaction with the NPC (Görlich et al., 1995; Moroianu et al., 1995) and carries importin-α with the NLS substrate into the nucleus. The translocation into the nucleus is terminated by the disassembly of the importin complex, and both subunits are returned probably separately to the cytoplasm. Importin-α interacts with -β via its importin-β binding domain (IBB domain; Görlich et al., 1996a ; Weis et al., 1996a ). Binding to importin-β with an IBB domain is sufficient for nuclear entry, and the IBB domain can therefore be regarded as the nuclear targeting signal of importin-α. The export domain of importin-α has not yet been identified, but it is distinct from the IBB domain.The small GTPase Ran (Drivas et al., 1990; Bischoff and Ponstingl, 1991b ; Belhumeur et al., 1993) plays a key role in NLS-dependent protein import (Melchior et al., 1993; Moore and Blobel, 1993). GTP hydrolysis by Ran is absolutely essential for import (Melchior et al., 1993; Moore and Blobel, 1993; Schlenstedt et al., 1995a ; Palacios et al., 1996) and is possibly even its sole source of energy (Weis et al., 1996b ). Although the molecular mechanism of import is far from being understood, it appears that Ran fulfils at least two distinct functions in this process: first, Ran''s GTP cycle probably drives translocation into the nucleus (Melchior et al., 1993; Moore and Blobel, 1993; Weis et al., 1996b ), which appears to involve the binding of (cytoplasmic) RanGDP to the NPC, followed by nucleotide exchange and GTP hydrolysis, but it does not involve binding of RanGTP to importin-β (Görlich et al., 1996b ). Unfortunately, nothing is known of how Ran''s GTP cycle would translate into a directed movement through the NPC. Secondly, Ran regulates the interaction between importin-α and -β (Rexach and Blobel, 1995; Chi et al., 1996; Görlich et al., 1996b ). Binding of RanGTP to importin-β disassembles the importin-α/β complex at the nuclear side of the NPC, thereby terminating translocation (Görlich et al., 1996b ). The asymmetric distribution of Ran''s principal GDP/GTP exchange factor (RCC1; Bischoff and Ponstingl, 1991a ) and GTPase activating protein (RanGAP1, or RNA1 in yeast; Bischoff et al., 1995a ; Becker et al., 1995) crucially determines where the importin heterodimer can form and where it is forced to dissociate. RCC1 is a nuclear, chromatin-bound protein (Ohtsubo et al., 1987, 1989) that generates RanGTP in the nucleus, whereas free RanGTP is depleted from the cytoplasm by RanGAP1, which is excluded from the nucleoplasm (Hopper et al., 1990; Matunis et al., 1996; Mahajan et. al, 1997). Thus, low RanGTP levels in the cytoplasm allow importin-α to bind -β, and the high RanGTP concentration in the nuclear compartment dissociates the importin complex. The concentration of free RanGTP can, in this model, be regarded as a marker for cytoplasmic identity (low RanGTP) and nuclear identity (high RanGTP), which is “sensed” by the Ran-binding site in importin-β.It is likely that at least some properties of importin-β are shared by the mediators of the other nucleocytoplasmic transport pathways. This is emphasized by the recent identification of the importin-β–related transportin (Pollard et al., 1996) as an import receptor recognizing the M9 domain, the nuclear targeting signal in hnRNP A1 (Michael et al., 1995), and of yeast transportin (Kap 104p) as an import receptor for mRNA binding proteins (Aitchison et al., 1996). Furthermore, importin-β or its NPC-binding domain cross-compete with other pathways, such as M9-dependent import, NES-mediated nuclear export, and the export of mRNA and U snRNA (Kutay et al., 1997). This would suggest that these other transport receptors share at least some binding sites at the NPC and take a similar path through the nuclear pore complex as importin-β.In addition to importin-β, a number of other Ran-binding proteins are detectable in eukaryotic cells, e.g., in overlay blots using Ran γ-[32P]GTP as a probe. These can be grouped into two classes (Lounsbury et al., 1994, 1996): first, those with a RanBP1 homology domain including the Ran binding protein 1 (RanBP1) itself (Coutavas et al., 1993; Bischoff et al., 1995b ) and the nuclear pore protein RanBP2, which has four RanBP1 homology domains (Wu et al., 1995; Yokoyama et al., 1995). Their binding to Ran can be competed by RanBP1. Second, importin-β and so far unidentified protein(s) of ∼120 kD whose Ran-binding is competed by importin-β but not by excess of RanBP1 (Lounsbury et al., 1994, 1996). Both RanBP1 and importin-β inhibit the nucleotide exchange on RanGTP (Coutavas et al., 1993; Lounsbury et al., 1994, 1996; Bischoff et al., 1995b ; Görlich et al., 1996b ). However, they do not cross-compete with each other for Ran binding but instead bind to different, nonoverlapping sites on Ran (Chi et al., 1996; Kutay et al., 1997; Lounsbury and Macara, 1997). Another striking difference is that RanBP1 facilitates the activation of Ran''s GTPase by RanGAP1 (Beddow et al., 1995; Bischoff et al., 1995b ), whereas the importin-β/RanGTP complex is entirely GAP resistant (Floer and Blobel, 1996; Görlich et al., 1996b ).Although a direct involvement of Ran has so far only been demonstrated in the importin-dependent transport pathway, perturbations in the Ran system have severe effects on a great variety of cellular functions, such as RNA processing, RNA export, regulation of chromosome structure, cell cycle progression, initiation of replication, microtubule structure, etc. (for review see Dasso, 1993; Sazer, 1996). One could argue that these effects are all secondary consequences from an impaired NLS-dependent protein import. However, it is also possible that these defects are more direct and that eukaryotic cells contain many immediate targets of Ran function.Here we describe a novel superfamily of Ran-binding proteins, which includes about a dozen factors in yeast and probably even more in higher eukaryotes. The members of this superfamily share with importin-β an NH2-terminal sequence motif that appears to account for RanGTP binding. Indeed we could confirm the interaction with Ran for the following factors: RanBP7 and RanBP8, two novel, related proteins described here, Cse1p, a cell cycle regulator in yeast, CAS, which is required for apoptosis in cultured human cells, and for Msn5p and Pse1p from yeast. Of these we have characterized RanBP7 and RanBP8 in more detail. Both resemble closely importin-β in their interaction with Ran, and both bind directly to nuclear pore complexes. RanBP7 can cross the nuclear membrane rapidly and in both directions. We provide evidence for a transport cycle in which RanBP7 first enters the nucleus, binds RanGTP inside the nucleus as a prerequisite for rapid re-export to the cytoplasm, after which the RanBP7/RanGTP complex becomes finally disassembled by the concerted action of RanBP1 and RanGAP1 in the cytoplasm. We propose that during these transport cycles, RanBP7 would normally carry an as yet unidentified cargo. This means, RanBP7 and possibly also the other members of the RanBP7/Cse1p/ importin-β superfamily could function as transport receptors that shuttle between nucleus and cytoplasm. RanBP7 and importin-β form an abundant, heterodimeric complex in the cytoplasm that appears to have a function different from nuclear import of proteins with a classical NLS. It might be a way to regulate either RanBP7 or importin-β function. Alternatively, the RanBP7/importin-β complex might constitute an import receptor with a substrate specificity different from that of the importin-α/β complex.  相似文献   

17.
SEC35 was identified in a novel screen for temperature-sensitive mutants in the secretory pathway of the yeast Saccharomyces cerevisiae (Wuestehube et al., 1996. Genetics. 142:393–406). At the restrictive temperature, the sec35-1 strain exhibits a transport block between the ER and the Golgi apparatus and accumulates numerous vesicles. SEC35 encodes a novel cytosolic protein of 32 kD, peripherally associated with membranes. The temperature-sensitive phenotype of sec35-1 is efficiently suppressed by YPT1, which encodes the rab-like GTPase required early in the secretory pathway, or by SLY1-20, which encodes a dominant form of the ER to Golgi target -SNARE–associated protein Sly1p. Weaker suppression is evident upon overexpression of genes encoding the vesicle-SNAREs SEC22, BET1, or YKT6. The cold-sensitive lethality that results from deleting SEC35 is suppressed by YPT1 or SLY1-20. These genetic relationships suggest that Sec35p acts upstream of, or in conjunction with, Ypt1p and Sly1p as was previously found for Uso1p. Using a cell-free assay that measures distinct steps in vesicle transport from the ER to the Golgi, we find Sec35p is required for a vesicle docking stage catalyzed by Uso1p. These genetic and biochemical results suggest Sec35p acts with Uso1p to dock ER-derived vesicles to the Golgi complex.Protein transport through the secretory pathway occurs via transport vesicles under the direction of a large set of protein components (Rothman, 1994). The process can be divided into three stages: (a) vesicle budding, (b) vesicle docking, and (c) membrane fusion, with distinct sets of proteins mediating each phase. The budding step involves recruitment of coat proteins to the membrane and culminates with the release of coated vesicles (Schekman and Orci, 1996). The docking reaction is likely to require a set of integral membrane proteins on the vesicle and target membranes, termed v-SNAREs1 and t-SNAREs (vesicle- and target membrane-soluble N-ethylmaleimide–sensitive fusion protein [NSF] attachment protein [SNAP] receptors, respectively), that are thought to confer specificity through their pair-wise interactions (Söllner et al., 1993b ). Small GTP-binding proteins of the rab family also assist in the docking process (Ferro-Novick and Novick, 1993), but their precise function is not known. The fusion step ensues after docking and results in the delivery of the vesicular cargo to the next compartment in the secretory pathway.Vesicular transport from the ER to the Golgi apparatus in the yeast Saccharomyces cerevisiae has been extensively characterized. Transport vesicle budding involves the assembly of the COPII coat, composed of the Sec13p/Sec31p (Pryer et al., 1993; Salama et al., 1993; Barlowe et al., 1994) and Sec23p/Sec24p heterodimers (Hicke and Schekman, 1989; Hicke et al., 1992), under the direction of an integral membrane protein, Sec12p (Nakano et al., 1988; Barlowe and Schekman, 1993), a small GTPase, Sar1p (Nakano and Muramatsu, 1989), and a multidomain protein, Sec16p (Espenshade et al., 1995; Shaywitz et al., 1997). Docking is thought to require a tethering event mediated by Uso1p (Cao et al., 1998), the yeast homologue of mammalian p115 (Barroso et al., 1995; Sapperstein et al., 1995), followed by or concurrent with the interaction of a set of ER to Golgi v-SNAREs, Bet1p, Bos1p, Sec22p (Newman and Ferro-Novick, 1987; Newman et al., 1990; Ossig et al., 1991; Shim et al., 1991; Søgaard et al., 1994) and perhaps Ykt6p (Søgaard et al., 1994; McNew et al., 1997), with the cognate t-SNARE on the Golgi, Sed5p (Hardwick and Pelham, 1992). For some time it was thought that fusion may be initiated by disassembly of the v/t-SNARE complex (Söllner et al., 1993a ) by yeast SNAP, Sec17p, (Griff et al., 1992) and NSF, Sec18p (Eakle et al., 1988; Wilson et al., 1989). However, this concept has been challenged by studies with a yeast system that reconstitutes homotypic vacuolar fusion, which suggests the action of Sec18p is before vesicle docking (Mayer et al., 1996; Mayer and Wickner, 1997). In addition, a prefusion role for NSF has been supported by the recent finding that liposomes bearing SNAREs alone can fuse in the absence of NSF (Weber et al., 1998).Several proteins involved in the regulation of yeast ER to Golgi v/t-SNARE complex assembly have been identified, including Ypt1p, Uso1p, and Sly1p. Ypt1p is a member of the rab family of small GTP-binding proteins that have been identified as important components of almost every stage in the secretory pathway (Ferro-Novick and Novick, 1993). Hydrolysis of GTP by rab-like proteins has been hypothesized to provide the regulatory switch that controls the fidelity of vesicular transport (Bourne, 1988). A second protein, Uso1p (Nakajima et al., 1991), appears to function in the same pathway as Ypt1p (Sapperstein et al., 1996), and both proteins have been demonstrated to be essential for SNARE complex assembly (Søgaard et al., 1994; Sapperstein et al., 1996; Lupashin and Waters, 1997). The third protein, Sly1p, is associated with the t-SNARE Sed5p (Søgaard et al., 1994). SLY1 is an essential gene in yeast (Dascher et al., 1991; Ossig et al., 1991), and Sly1p is required for ER to Golgi transport in vitro (Lupashin et al., 1996) and in vivo (Ossig et al., 1991). However, several lines of evidence, particularly from Sly1p homologues in other organisms, indicate that Sly1p may also function as a negative regulator of v/t-SNARE complex assembly, perhaps by preventing the association of the v- and t-SNAREs (Hosono et al., 1992; Pevsner et al., 1994; Schulze et al., 1994). A dominant allele of SLY1, termed SLY1-20, is capable of suppressing mutations in YPT1 and USO1, including complete deletions (Dascher et al., 1991; Sapperstein et al., 1996). Thus, in the presence of Sly1-20p, two components required for SNARE complex assembly are no longer essential. We have proposed a model (Sapperstein et al., 1996; Lupashin and Waters, 1997) in which Ypt1p and Uso1p function to relieve the inhibitory action of Sly1p on SNARE complex assembly. In this model Sly1-20p can be thought of as a noninhibitory form of SLY1 that renders Ypt1p and Uso1p superfluous.We believe that the ability of SLY1-20 to suppress defects in upstream docking regulators can be used to identify additional components involved in the regulation of vesicular docking. We have undertaken a genetic screen (to be presented elsewhere) to isolate novel components in this pathway which, when mutated, depend upon Sly1-20p for viability. In the course of this work, we discovered that two recently identified mutants, sec34 and sec35, can be suppressed by SLY1-20 and thus satisfy the criterion of our screen. These mutants were isolated in a novel screen to identify components involved in transport at any step between the ER and the trans-Golgi network (i.e., the Kex2p compartment) in yeast (Wuestehube et al., 1996). Both sec34 and sec35 accumulate the core-glycosylated form of secretory proteins at the nonpermissive temperature, indicating a block in ER to Golgi transport. Furthermore, electron microscopy indicated that both sec34 and sec35 accumulate numerous vesicles upon shift to the restrictive temperature (Wuestehube et al., 1996), a hallmark of genes whose protein products are involved in the docking or fusion phase of transport (Kaiser and Schekman, 1990). In this report we describe the cloning of SEC35 and analysis of its genetic interactions with other secretory genes. Strong genetic interaction between SEC35 and SLY1, YPT1, and USO1 suggests that Sec35p may function in vesicle docking. To test this possibility, we devised an in vitro transport assay that depends on the addition of purified Sec35p and Uso1p. Vesicles synthesized in the absence of functional Sec35p do not fuse with the Golgi compartment and remain as freely diffusible intermediates. Upon addition of Sec35p and Uso1p, vesicles dock to the Golgi and proceed to membrane fusion. Requirements for Sec35p at the vesicle docking step correlates our genetic experiments with the biochemically distinguishable steps of vesicle docking and membrane fusion.  相似文献   

18.
19.
This year marks the 35th anniversary of the isolation of 23 SEC genes. These genes all encode key regulators of the secretory pathway, and much of our knowledge of the secretory pathway is based on this initial discovery. The identification of the SEC genes is a result of combining genetics, biochemistry, and electron microscopy in a very clever way. Scientists have been busy ever since seeking to understand the function and regulation of these genes and to identify further key players in the process. Although most of the machinery acting along the secretory pathway is known and its function generally understood, knowledge of regulation of the pathway under various conditions is still scarce and will keep researchers busy for years to come.Scientists have been fascinated by cells since their discovery in the 17th century—how they function and how they communicate with their neighbors and environment. However, research efforts were hindered for a long time by the lack of suitable tools with which to interrogate cellular function. Then in the 1930s and 1940s transmission electron microscopy (TEM) was invented and developed, which paved the way to investigating the structure of cellular organelles through the imaging of thin sections. TEM permitted for the first time a view of cells at low-nanometer resolution, well beyond the resolution of light microscopy (∼0.2 μm). At the time, this technological advance generated probably as much hype as there has been in recent years for superresolution light microscopy techniques.In spite of the ability to see ribosomes, the endoplasmic reticulum (ER), the Golgi apparatus, lysosomes, and vesicles in thin sections, only limited information could be gained about the dynamics of the secretory pathway. George Palade pioneered a technique in which he combined pulse-chase labeling with electron microscopy, revealing how proteins were transported along the secretory pathway (Jamieson and Palade, 1968 , 1971 ). However, no connection was available linking observable static TEM images to the identity of the players that would enable the cell to form organelles and promote the communication between them.In parallel, another powerful approach to understanding cellular function was the development of cell fractionation and biochemical assays. These methods allowed the identification of cellular processes and their reconstitution in vitro. Major insights were gained into mitochondrial function, DNA replication, RNA and protein synthesis, and lipid biosynthesis, just to name a few. Nonetheless, although proteins responsible for certain enzymatic activities could be purified, the identity of most cellular players remained elusive in the pregenomic era.The third pillar for the interrogation of cellular functions came through genetics. In the late 1960s to early 1970s, genetic screens for conditional, temperature-sensitive mutants revealed mutants that would arrest at particular points in the cell cycle (Hartwell, 1967 ; Hartwell et al., 1970 ; Nurse, 1975 ; Nurse et al., 1976 ).This was around the time that Randy Schekman set up his lab at the University of California in Berkeley. Schekman had been trained as a biochemist working on DNA replication but became interested in membranes, organelles, and secretion, inspired by work of George Palade, Albert Claude, and Christian De Duve. It was also still a time when a young assistant professor did not necessarily have to continue on and further develop the research theme from his previous experience but could start something completely new. However, already back then, the National Institutes of Health refused to fund Schekman''s first grant application because he had no experience working with yeast and had no preliminary data. When Palade visited Berkeley and met Schekman soon after they had initiated the project, Palade was surprised to learn that yeast cells make glycoproteins.Fascinated by the success of genetic screens for conditional mutants in yeast, Schekman, together with his first graduate student, Peter Novick, set out to screen for temperature-sensitive mutants that would be defective in secretion. As so often happens, the first, initial and clever idea for the screen did not work out. However, Schekman and Novick were not so easily discouraged and came up with another idea. They took temperature-sensitive mutants from the initial screen and measured the accumulation of secretory enzymes in cells using simple colorimetric assays of invertase and acid phosphatase activity (Novick and Schekman, 1979 ). When they looked more carefully, by TEM, at some mutants with defects in enzyme secretion, they realized the mutant cells were full of vesicles and accumulated internal membranes. The very first mutant, sec1-1, was found to block the fusion of transport vesicles with the plasma membrane (Novick and Schekman, 1979 ). From this observation that secretion mutants accumulated proteins and membranes inside the cell (Novick and Schekman, 1979 ) they suspected that mutant yeast cells would possess different physicochemical properties than wild-type cells. Indeed, Novick could show that mutant cells displayed an increase in buoyant density. The difference provided a powerful screening procedure through which secretory mutants could be enriched by gradient centrifugation while complementation groups were determined by classical genetics. In 1980, Novick, Field, and Schekman published the landmark paper in which they reported the discovery of 23 complementation groups involved in posttranslational events in the secretory pathway in Saccharomyces cerevisiae (Novick et al., 1980 ). Within 1 year, Novick, Ferro, and Schekman managed to assign the 23 genes to specific steps along the secretory pathway (Novick et al., 1981 ; Figure 1). Not only had they discovered 23 genes involved in secretion, but they had also elegantly combined three very prominent investigation tools: electron microscopy, biochemistry, and genetics. This powerful mix of approaches enabled Schekman and many coworkers over the years to craft an impressive picture describing the path of a nascent secretory protein from when it leaves the ribosome and enters the ER until its discharge at the plasma membrane. The realization in the 1980s and 1990s that the basic transport machineries are conserved from yeast to human made yeast THE model of choice to identify genes involved not only in secretion but also in endocytosis and lysosomal/vacuolar sorting pathways (Riezman, 1985 ; Bankaitis et al., 1986 ; Chvatchko et al., 1986 ; Rothman and Stevens, 1986 ; Rothman et al., 1989 ; Robinson et al., 1988 ).Open in a separate windowFIGURE 1:Yeast secretory pathway. BB, Berkeley body; CW, cell wall; ER, endoplasmic reticulum; PM, plasma membrane; SEC, wild-type gene product; sec, mutant gene product. Asterisks denote gene products that were found later to function along the entire secretory pathway, in fact, at all membrane fusion steps. Redrawn from Novick et al. (1981) .After the identification of the mutants and cloning of the genes, Schekman went back to his roots in biochemistry, in keeping with the immortal words of Richard Feynman, “what I cannot create, I do not understand.” The mantra in the lab for years to come was reconstitution— the in vitro establishment of transport processes such as polypeptide translocation into the ER and formation of transport vesicles. Schekman''s lab over the years came up with a number of in vitro systems that faithfully recapitulated transport processes and greatly influenced our current understanding of the secretory pathway.What happened to the brave graduate student Peter Novick? He was hooked on the secretory pathway: after his postdoc on yeast actin, he went on to discover, among other things, a Rab GTPase and its activator, the exocyst complex, and the soluble N-ethylmaleimide–sensitive factor attachment protein receptors, promoting fusion of secretory vesicles with the plasma membrane (Salminen and Novick, 1987 ; Goud et al., 1988 ; Bowser and Novick, 1991 ; Brennwald et al., 1994 ; TerBush and Novick, 1995 ; Walch-Solimena et al., 1997 ). Most of theses components were among the 23 gene products identified in the original screen, and the mutations blocked fusion of secretory vesicles with the plasma membrane (Figure 1).The functions of all 23 gene products that Novick and Schekman initially identified are now generally understood, and crystal structures are available for many of them, providing insights into the basic machineries that operate along the secretory pathway. In spite of all this knowledge, we are only starting to understand the regulation and fine-tuning of the secretory pathway to ensure proper protein and lipid distribution and to maintain protein and lipid homeostasis upon aging, stress, and environmental changes.It all started with this sec1-1 mutant filled with secretory vesicles some 35 years ago, and it will take us perhaps the same amount of time to fully understand the regulation of intracellular transport pathways.  相似文献   

20.
We purified from rat brain a novel actin filament (F-actin)–binding protein of ∼180 kD (p180), which was specifically expressed in neural tissue. We named p180 neurabin (neural tissue–specific F-actin– binding protein). We moreover cloned the cDNA of neurabin from a rat brain cDNA library and characterized native and recombinant proteins. Neurabin was a protein of 1,095 amino acids with a calculated molecular mass of 122,729. Neurabin had one F-actin–binding domain at the NH2-terminal region, one PSD-95, DlgA, ZO-1–like domain at the middle region, a domain known to interact with transmembrane proteins, and domains predicted to form coiled-coil structures at the COOH-terminal region. Neurabin bound along the sides of F-actin and showed F-actin–cross-linking activity. Immunofluorescence microscopic analysis revealed that neurabin was highly concentrated in the synapse of the developed neurons. Neurabin was also concentrated in the lamellipodia of the growth cone during the development of neurons. Moreover, a study on suppression of endogenous neurabin in primary cultured rat hippocampal neurons by treatment with an antisense oligonucleotide showed that neurabin was involved in the neurite formation. Neurabin is a candidate for key molecules in the synapse formation and function.During the development of the nervous system, the distal tip of the elongating axon—the growth cone—actively migrates toward its target cell in response to the combined actions of attractive and repulsive guidance molecules in the extracellular environment (Garrity and Zipursky, 1995; Keynes and Cook, 1995; Chiba and Keshishian, 1996; Culotti and Kolodkin, 1996; Friedman and O''Leary, 1996; Tessier-Lavigne and Goodman, 1996). When the growth cone contacts with the target cell, it is transformed into the functional presynaptic terminal (Garrity and Zipursky, 1995; Chiba and Kishishian, 1996). The actin cytoskeleton has been shown to play crucial roles in these processes of the synapse formation (Mitchison and Kirschner, 1988; Smith, 1988; Bentley and O''Connor, 1994; Lin et al., 1994; Mackay et al., 1995; Tanaka and Sabry, 1995).In the developing nervous system, the actin cytoskeleton is prominent in two structural domains of the growth cone, filopodia and lamellipodia (Mitchison and Kirschner, 1988; Smith, 1988; Bentley and O''Connor, 1994; Lin et al., 1994; Mackay et al., 1995; Tanaka and Sabry, 1995). In these domains, actin filament (F-actin)1 assembled at the leading edge are transported into the center of the growth cone and disassembled there. It has been suggested that this retrograde flow of F-actin is crucial for the growth cone motility. Drugs that disrupt F-actin have also been shown to cause the lamellipodial and filopodial collapse and block the ability of neurons to extend the growth cone in the correct direction (Marsh and Letourneau, 1984; Forscher and Smith, 1988; Bentley and Toroian-Raymond, 1986; Chien et al., 1993). These results suggest that the actin cytoskeleton regulates not only the growth cone motility but also the growth cone directionality. Recently, a variety of guidance molecules and their receptors have been identified (Garrity and Zipursky, 1995; Keynes and Cook, 1995; Chiba and Keshishian, 1996; Culotti and Kolodkin, 1996; Friedman and O''Leary, 1996; Tessier-Lavigne and Goodman, 1996). However, which molecules of the actin cytoskeleton are essential for the growth cone motility and directionality is not well understood.When the growth cone contacts with the target cell, the target cell regulates the development of the presynaptic nerve terminal and the formation of the functional synapse (Bowe and Fallon, 1995; Chiba and Keshishian, 1996). In the established nervous system, the presynaptic and postsynaptic membranes get aligned in space and constitute the synaptic junction (Burns and Augustine, 1995; Garner and Kindler, 1996). Electron microscopic studies have revealed the ultrastructural features of the synaptic junction (Burns and Augustine, 1995; Garner and Kindler, 1996). The presynaptic cytoplasm is characterized by synaptic vesicles (SVs). SVs are not distributed uniformly; SVs cluster together in the vicinity of the presynaptic plasma membrane, where F-actin forms a network and is associated with the presynaptic plasma membrane (Hirokawa et al., 1989). Most SVs within the cluster are linked through thin strands to each other, to F-actin, or to both (Hirokawa et al., 1989). A subset of SVs within the cluster are attached by fine filamentous threads to neurotransmitter release zone at the presynaptic plasma membrane (Hirokawa et al., 1989). The presynaptic submembranous cytoskeleton is assumed to be involved in recruiting Ca2+ channels and the components of the SV fusion complex, delivering SVs to the neurotransmitter release zone, and keeping them in place (Burns and Augustine, 1995; Garner and Kindler, 1996). At the inner surface of the post-synaptic plasma membrane, there is an electron dense thickening, called postsynaptic density. The postsynaptic density is assumed to be involved in the selective targeting and accumulation of ion channels and receptors (Burns and Augustine, 1995; Garner and Kindler, 1996). It is also assumed that the presynaptic and postsynaptic submembranous cytoskeleton elements are linked to cell adhesion molecules to regulate the synaptic stabilization and plasticity (Fields and Itoh, 1996; Garner and Kindler, 1996). The presynaptic and postsynaptic submembranous cytoskeleton elements are thought to be composed of spectrin/fodrin, ankyrin, α-adducin, and protein 4.1 isoforms and to be linked to F-actin through these cytoskeleton proteins (Garner and Kindler, 1996). However, little is known about which molecules of the submembranous cytoskeleton are essential for the synaptic transmission and/or the synaptic stabilization.To understand the regulation of the actin cytoskeleton during and after the development of the nervous system, it is of crucial importance to identify F-actin–binding proteins implicated in the synapse formation and function. Therefore, we attempted here to isolate neural tissue–specific F-actin–binding proteins. We isolated a novel neural tissue–specific F-actin–binding protein from rat brain, which may be involved in neurite formation, and named it neurabin (neural tissue–specific F-actin–binding protein).  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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