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 31 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.
In vitro and in vivo studies implicate occludin in the regulation of paracellular macromolecular flux at steady state and in response to tumor necrosis factor (TNF). To define the roles of occludin in these processes, we established intestinal epithelia with stable occludin knockdown. Knockdown monolayers had markedly enhanced tight junction permeability to large molecules that could be modeled by size-selective channels with radii of ∼62.5 Å. TNF increased paracellular flux of large molecules in occludin-sufficient, but not occludin-deficient, monolayers. Complementation using full-length or C-terminal coiled-coil occludin/ELL domain (OCEL)–deficient enhanced green fluorescent protein (EGFP)–occludin showed that TNF-induced occludin endocytosis and barrier regulation both required the OCEL domain. Either TNF treatment or OCEL deletion accelerated EGFP-occludin fluorescence recovery after photobleaching, but TNF treatment did not affect behavior of EGFP-occludinΔOCEL. Further, the free OCEL domain prevented TNF-induced acceleration of occludin fluorescence recovery, occludin endocytosis, and barrier loss. OCEL mutated within a recently proposed ZO-1–binding domain (K433) could not inhibit TNF effects, but OCEL mutated within the ZO-1 SH3-GuK–binding region (K485/K488) remained functional. We conclude that OCEL-mediated occludin interactions are essential for limiting paracellular macromolecular flux. Moreover, our data implicate interactions mediated by the OCEL K433 region as an effector of TNF-induced barrier regulation.Tight junctions seal the paracellular space in simple epithelia, such as those lining the lungs, intestines, and kidneys (Anderson et al., 2004 ; Fanning and Anderson, 2009 ; Shen et al., 2011 ). In the intestine, reduced paracellular barrier function is associated with disorders in which increased paracellular flux of ions and molecules contributes to symptoms such as diarrhea, malabsorption, and intestinal protein loss. Recombinant tumor necrosis factor (TNF) can be used to model this barrier loss in vitro or in vivo (Taylor et al., 1998 ; Clayburgh et al., 2006 ), and TNF neutralization is associated with restoration of intestinal barrier function in Crohn''s disease (Suenaert et al., 2002 ). Further, in vivo and in vitro studies of intestinal epithelia show that TNF-induced barrier loss requires myosin light chain kinase (MLCK) activation (Zolotarevsky et al., 2002 ; Clayburgh et al., 2005 , 2006 ; Ma et al., 2005 ; Wang et al., 2005 ). The resulting myosin II regulatory light chain (MLC) phosphorylation drives occludin internalization, which is required for cytokine-induced intestinal epithelial barrier loss (Clayburgh et al., 2005 , 2006 ; Schwarz et al., 2007 ; Marchiando et al., 2010 ). In addition, transgenic EGFP-occludin expression in vivo limits TNF-induced depletion of tight junction–associated occludin, barrier loss, and diarrhea (Marchiando et al., 2010 ). Conversely, in vitro studies show that occludin knockdown limits TNF-induced barrier regulation (Van Itallie et al., 2010 ). The basis for this discrepancy is not understood.One challenge is that, despite being identified 20 yr ago (Furuse et al., 1993 ), the contribution of occludin to tight junction regulation remains incompletely defined. The observation that occludin-knockout mice are able to form paracellular barriers and do not have obvious defects in epidermal, respiratory, or bladder tight junction function (Saitou et al., 2000 ; Schulzke et al., 2005 ) led many to conclude that occludin is not essential for tight junction barrier function. It is important to note, however, that barrier regulation in response to stress has not been studied in occludin-deficient animals.We recently showed that dephosphorylation of occludin serine-408 promotes assembly of a complex composed of occludin, ZO-1, and claudin-2 that inhibits flux across size- and charge-selective channels termed the pore pathway (Anderson and Van Itallie, 2009 ; Turner, 2009 ; Raleigh et al., 2011 ; Shen et al., 2011 ). Although this demonstrates that occludin can serve a regulatory role, it does not explain the role of occludin in TNF-induced barrier loss, which increases flux across the size- and charge-nonselective leak pathway (Wang et al., 2005 ; Weber et al., 2010 ). In vitro studies, however, do suggest that occludin contributes to leak pathway regulation, as occludin knockdown in either Madin–Darby canine kidney (MDCK) or human intestinal (Caco-2) epithelial monolayers enhances leak pathway permeability (Yu et al., 2005 ; Al-Sadi et al., 2011 ; Ye et al., 2011 ). Taken as a whole, these data suggest that occludin organizes the tight junction to limit leak pathway flux, whereas occludin removal, either by knockdown or endocytosis, enhances leak pathway flux.To define the mechanisms by which occludin regulates the leak pathway, we analyzed the contributions of occludin, as well as specific occludin domains, to basal and TNF-induced barrier regulation. The data indicate that TNF destabilizes tight junction–associated occludin via interactions mediated by the C-terminal coiled-coil occludin/ELL domain (OCEL). Further, these OCEL-mediated events are required for TNF-induced barrier regulation. Thus these data provide new insight into the structural elements and mechanisms by which occludin regulates leak pathway paracellular flux.  相似文献   

3.
4.
5.
Leptospira spp., the causative agents of leptospirosis, adhere to components of the extracellular matrix, a pivotal role for colonization of host tissues during infection. Previously, we and others have shown that Leptospira immunoglobulin-like proteins (Lig) of Leptospira spp. bind to fibronectin, laminin, collagen, and fibrinogen. In this study, we report that Leptospira can be immobilized by human tropoelastin (HTE) or elastin from different tissues, including lung, skin, and blood vessels, and that Lig proteins can bind to HTE or elastin. Moreover, both elastin and HTE bind to the same LigB immunoglobulin-like domains, including LigBCon4, LigBCen7′–8, LigBCen9, and LigBCen12 as demonstrated by enzyme-linked immunosorbent assay (ELISA) and competition ELISAs. The LigB immunoglobulin-like domain binds to the 17th to 27th exons of HTE (17–27HTE) as determined by ELISA (LigBCon4, KD = 0.50 μm; LigBCen7′–8, KD = 0.82 μm; LigBCen9, KD = 1.54 μm; and LigBCen12, KD = 0.73 μm). The interaction of LigBCon4 and 17–27HTE was further confirmed by steady state fluorescence spectroscopy (KD = 0.49 μm) and ITC (KD = 0.54 μm). Furthermore, the binding was enthalpy-driven and affected by environmental pH, indicating it is a charge-charge interaction. The binding affinity of LigBCon4D341N to 17–27HTE was 4.6-fold less than that of wild type LigBCon4. In summary, we show that Lig proteins of Leptospira spp. interact with elastin and HTE, and we conclude this interaction may contribute to Leptospira adhesion to host tissues during infection.Pathogenic Leptospira spp. are spirochetes that cause leptospirosis, a serious infectious disease of people and animals (1, 2). Weil syndrome, the severe form of leptospiral infection, leads to multiorgan damage, including liver failure (jaundice), renal failure (nephritis), pulmonary hemorrhage, meningitis, abortion, and uveitis (3, 4). Furthermore, this disease is not only prevalent in many developing countries, it is reemerging in the United States (3). Although leptospirosis is a serious worldwide zoonotic disease, the pathogenic mechanisms of Leptospira infection remain enigmatic. Recent breakthroughs in applying genetic tools to Leptospira may facilitate studies on the molecular pathogenesis of leptospirosis (58).The attachment of pathogenic Leptospira spp. to host tissues is critical in the early phase of Leptospira infection. Leptospira spp. adhere to host tissues to overcome mechanical defense systems at tissue surfaces and to initiate colonization of specific tissues, such as the lung, kidney, and liver. Leptospira invade hosts tissues through mucous membranes or injured epidermis, coming in contact with subepithelial tissues. Here, certain bacterial outer surface proteins serve as microbial surface components recognizing adhesive matrix molecules (MSCRAMMs)2 to mediate the binding of bacteria to different extracellular matrices (ECMs) of host cells (9). Several leptospiral MSCRAMMs have been identified (1018), and we speculate that more will be identified in the near future.Lig proteins are distributed on the outer surface of pathogenic Leptospira, and the expression of Lig protein is only found in low passage strains (14, 16, 17), probably induced by environmental cues such as osmotic or temperature changes (19). Lig proteins can bind to fibrinogen and a variety of ECMs, including fibronectin (Fn), laminin, and collagen, thereby mediating adhesion to host cells (2023). Lig proteins also constitute good vaccine candidates (2426).Elastin is a component of ECM critical to tissue elasticity and resilience and is abundant in skin, lung, blood vessels, placenta, uterus, and other tissues (2729). Tropoelastin is the soluble precursor of elastin (28). During the major phase of elastogenesis, multiple tropoelastin molecules associate through coacervation (3032). Because of the abundance of elastin or tropoelastin on the surface of host cells, several bacterial MSCRAMMs use elastin and/or tropoelastin to mediate adhesion during the infection process (3335).Because leptospiral infection is known to cause severe pulmonary hemorrhage (36, 37) and abortion (38), we hypothesize that some leptospiral MSCRAMMs may interact with elastin and/or tropoelastin in these elastin-rich tissues. This is the first report that Lig proteins of Leptospira interact with elastin and tropoelastin, and the interactions are mediated by several specific immunoglobulin-like domains of Lig proteins, including LigBCon4, LigBCen7′–8, LigBCen9, and LigBCen12, which bind to the 17th to 27th exons of human tropoelastin (HTE).  相似文献   

6.
7.
Sperm glyceraldehyde-3-phosphate dehydrogenase has been shown to be a successful target for a non-hormonal contraceptive approach, but the agents tested to date have had unacceptable side effects. Obtaining the structure of the sperm-specific isoform to allow rational inhibitor design has therefore been a goal for a number of years but has proved intractable because of the insoluble nature of both native and recombinant protein. We have obtained soluble recombinant sperm glyceraldehyde-3-phosphate dehydrogenase as a heterotetramer with the Escherichia coli glyceraldehyde-3-phosphate dehydrogenase in a ratio of 1:3 and have solved the structure of the heterotetramer which we believe represents a novel strategy for structure determination of an insoluble protein. A structure was also obtained where glyceraldehyde 3-phosphate binds in the Ps pocket in the active site of the sperm enzyme subunit in the presence of NAD. Modeling and comparison of the structures of human somatic and sperm-specific glyceraldehyde-3-phosphate dehydrogenase revealed few differences at the active site and hence rebut the long presumed structural specificity of 3-chlorolactaldehyde for the sperm isoform. The contraceptive activity of α-chlorohydrin and its apparent specificity for the sperm isoform in vivo are likely to be due to differences in metabolism to 3-chlorolactaldehyde in spermatozoa and somatic cells. However, further detailed analysis of the sperm glyceraldehyde-3-phosphate dehydrogenase structure revealed sites in the enzyme that do show significant difference compared with published somatic glyceraldehyde-3-phosphate dehydrogenase structures that could be exploited by structure-based drug design to identify leads for novel male contraceptives.Glyceraldehyde-3-phosphate dehydrogenase-S (GAPDS3 in rat; GAPDH2 in human) is the sperm-specific isoform of GAPDH (13) and the sole GAPDH enzyme in sperm. GAPDS is highly conserved between species showing 94% identity between rat and mouse and 87% identity between rat and human. Within a particular species, GAPDS also shows significant sequence identity to its GAPDH paralogue, 70, 70, and 68% for rat, mouse, and human, respectively. The most striking difference between GAPDS and GAPDH is the presence of an N-terminal polyproline region in GAPDS, which is 97 residues in rat (accession number AJ297631), 105 in mouse (3), and 72 in human (2). GAPDS is restricted to the principal piece of the sperm flagellum (1, 2, 4) where it is localized to the fibrous sheath (5), an association proposed to be mediated via the N-terminal polyproline extension.GAPDS first came to prominence as a contraceptive target during the 1970s (68). Investigations showed that treatment of sperm with α-chlorohydrin or a number of related compounds could inhibit GAPDS activity (911), sperm motility (913), and the fertilization of oocytes in vitro (14). The metabolite of these compounds, 3-chlorolactaldehyde (1517), selectively inhibited GAPDS, having no effect on the activity of somatic cell GAPDH (18, 19), providing the specificity required for a potential contraceptive. Questions surrounding these particular compounds were raised when a number of side effects were evident from in vivo trials (7, 2022); however, the design of small molecule inhibitors of GAPDS may provide a viable alternative. Its potential as a contraceptive target was supported by data from mice where GAPDS−/− males (23) were infertile because of defects in sperm motility.Glyceraldehyde-3-phosphate dehydrogenases are tetrameric enzymes that catalyze the oxidative phosphorylation of d-glyceraldehyde 3-phosphate (Glc-3-P) into 1,3-diphosphoglycerate in the presence of an NAD cofactor via a two-step chemical mechanism (24). The first models of substrate binding were proposed on the basis of crystal structures of the holoenzyme from lobster (25) and Bacillus stearothermophilus (26), and Moras and co-workers (25) identified two anion-binding sites postulated to correspond to those binding the C-3 phosphate group of d-Glc-3-P (Ps site) and the inorganic phosphate ion (Pi site).Structure-based design of small molecules to inhibit GAPDH is not unprecedented. GAPDH has been targeted from protozoan parasites (2730), as the bloodstream forms rely solely on glycolysis for energy production (31, 32). A number of mammalian GAPDH structures have also been solved, including rabbit muscle (33, 34), human liver (35), and human placenta (36); however, no structures are available for sperm-specific isoforms of this enzyme.Active heterotetramers of GAPDH between different species have been reported and biochemically characterized previously, both in ratios of 2:2 and 3:1 (3740). In this study we have successfully obtained crystals of rat recombinant GAPDS as a heterotetramer with Escherichia coli GAPDH in a 1:3 ratio. To understand the basis of inhibition of the sperm isoform by substrate analogue 3-chlorolactaldehyde, a metabolite of α-chlorohydrin, a structure was also determined in the presence of the substrate glyceraldehyde 3-phosphate. The sperm-specific structure was compared with the human placental GAPDH structure (PDB entry 1U8F; Ref. 36) to identify differences that may provide a target for the design of inhibitors specific to the GAPDS protein. The unique structural features identified offer potential candidates for further investigation as inhibitor targets.  相似文献   

8.
9.
10.
11.
Mathematical tools developed in the context of Shannon information theory were used to analyze the meaning of the BLOSUM score, which was split into three components termed as the BLOSUM spectrum (or BLOSpectrum). These relate respectively to the sequence convergence (the stochastic similarity of the two protein sequences), to the background frequency divergence (typicality of the amino acid probability distribution in each sequence), and to the target frequency divergence (compliance of the amino acid variations between the two sequences to the protein model implicit in the BLOCKS database). This treatment sharpens the protein sequence comparison, providing a rationale for the biological significance of the obtained score, and helps to identify weakly related sequences. Moreover, the BLOSpectrum can guide the choice of the most appropriate scoring matrix, tailoring it to the evolutionary divergence associated with the two sequences, or indicate if a compositionally adjusted matrix could perform better.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]  相似文献   

12.
13.
A decoding algorithm is tested that mechanistically models the progressive alignments that arise as the mRNA moves past the rRNA tail during translation elongation. Each of these alignments provides an opportunity for hybridization between the single-stranded, -terminal nucleotides of the 16S rRNA and the spatially accessible window of mRNA sequence, from which a free energy value can be calculated. Using this algorithm we show that a periodic, energetic pattern of frequency 1/3 is revealed. This periodic signal exists in the majority of coding regions of eubacterial genes, but not in the non-coding regions encoding the 16S and 23S rRNAs. Signal analysis reveals that the population of coding regions of each bacterial species has a mean phase that is correlated in a statistically significant way with species () content. These results suggest that the periodic signal could function as a synchronization signal for the maintenance of reading frame and that codon usage provides a mechanism for manipulation of signal phase.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

14.
15.
Human concentrative nucleoside transporter 3 (hCNT3) utilizes electrochemical gradients of both Na+ and H+ to accumulate pyrimidine and purine nucleosides within cells. We have employed radioisotope flux and electrophysiological techniques in combination with site-directed mutagenesis and heterologous expression in Xenopus oocytes to identify two conserved pore-lining glutamate residues (Glu-343 and Glu-519) with essential roles in hCNT3 Na+/nucleoside and H+/nucleoside cotransport. Mutation of Glu-343 and Glu-519 to aspartate, glutamine, and cysteine severely compromised hCNT3 transport function, and changes included altered nucleoside and cation activation kinetics (all mutants), loss or impairment of H+ dependence (all mutants), shift in Na+:nucleoside stoichiometry from 2:1 to 1:1 (E519C), complete loss of catalytic activity (E519Q) and, similar to the corresponding mutant in Na+-specific hCNT1, uncoupled Na+ currents (E343Q). Consistent with close-proximity integration of cation/solute-binding sites within a common cation/permeant translocation pore, mutation of Glu-343 and Glu-519 also altered hCNT3 nucleoside transport selectivity. Both residues were accessible to the external medium and inhibited by p-chloromercuribenzene sulfonate when converted to cysteine.Physiologic nucleosides and the majority of synthetic nucleoside analogs with antineoplastic and/or antiviral activity are hydrophilic molecules that require specialized plasma membrane nucleoside transporter (NT)3 proteins for transport into or out of cells (14). NT-mediated transport is required for nucleoside metabolism by salvage pathways and is a critical determinant of the pharmacologic actions of nucleoside drugs (36). By regulating adenosine availability to purinoreceptors, NTs also modulate a diverse array of physiological processes, including neurotransmission, immune responses, platelet aggregation, renal function, and coronary vasodilation (4, 6, 7). Two structurally unrelated NT families of integral membrane proteins exist in human and other mammalian cells and tissues as follows: the SLC28 concentrative nucleoside transporter (CNT) family and the SLC29 equilibrative nucleoside transporter (ENT) family (3, 4, 6, 8, 9). ENTs are normally present in most, possibly all, cell types (4, 6, 8). CNTs, in contrast, are found predominantly in intestinal and renal epithelia and other specialized cell types, where they have important roles in absorption, secretion, distribution, and elimination of nucleosides and nucleoside drugs (13, 5, 6, 9).The CNT protein family in humans is represented by three members, hCNT1, hCNT2, and hCNT3. Belonging to a CNT subfamily phylogenetically distinct from hCNT1/2, hCNT3 utilizes electrochemical gradients of both Na+ and H+ to accumulate a broad range of pyrimidine and purine nucleosides and nucleoside drugs within cells (10, 11). hCNT1 and hCNT2, in contrast, are Na+-specific and transport pyrimidine and purine nucleosides, respectively (1113). Together, hCNT1–3 account for the three major concentrative nucleoside transport processes of human and other mammalian cells. Nonmammalian members of the CNT protein family that have been characterized functionally include hfCNT, a second member of the CNT3 subfamily from the ancient marine prevertebrate the Pacific hagfish Eptatretus stouti (14), CeCNT3 from Caenorhabditis elegans (15), CaCNT from Candida albicans (16), and the bacterial nucleoside transporter NupC from Escherichia coli (17). hfCNT is Na+- but not H+-coupled, whereas CeCNT3, CaCNT, and NupC are exclusively H+-coupled. Na+:nucleoside coupling stoichiometries are 1:1 for hCNT1 and hCNT2 and 2:1 for hCNT3 and hfCNT3 (11, 14). H+:nucleoside coupling ratios for hCNT3 and CaCNT are 1:1 (11, 16).Although much progress has been made in molecular studies of ENT proteins (4, 6, 8), studies of structurally and functionally important regions and residues within the CNT protein family are still at an early stage. Topological investigations suggest that hCNT1–3 and other eukaryote CNT family members have a 13 (or possibly 15)-transmembrane helix (TM) architecture, and multiple alignments reveal strong sequence similarities within the C-terminal half of the proteins (18). Prokaryotic CNTs lack the first three TMs of their eukaryotic counterparts, and functional expression of N-terminally truncated human and rat CNT1 in Xenopus oocytes has established that these three TMs are not required for Na+-dependent uridine transport activity (18). Consistent with this finding, chimeric studies involving hCNT1 and hfCNT (14) and hCNT1 and hCNT3 (19) have demonstrated that residues involved in Na+- and H+-coupling reside in the C-terminal half of the protein. Present in this region of the transporter, but of unknown function, is a highly conserved (G/A)XKX3NEFVA(Y/M/F) motif common to all eukaryote and prokaryote CNTs.By virtue of their negative charge and consequent ability to interact directly with coupling cations and/or participate in cation-induced and other protein conformational transitions, glutamate and aspartate residues play key functional and structural roles in a broad spectrum of mammalian and bacterial cation-coupled transporters (2030). Little, however, is known about their role in CNTs. This study builds upon a recent mutagenesis study of conserved glutamate and aspartate residues in hCNT1 (31) to undertake a parallel in depth investigation of corresponding residues in hCNT3. By employing the multifunctional capability of hCNT3 as a template for these studies, this study provides novel mechanistic insights into the molecular mechanism(s) of CNT-mediated cation/nucleoside cotransport, including the role of the (G/A)XKX3NEFVA(Y/M/F) motif.  相似文献   

16.
Matrix metalloproteinases (MMPs) have been extensively studied because of their functional attributes in development and diseases. However, relatively few in vivo functional studies have been reported on the roles of MMPs in postembryonic organ development. Amphibian metamorphosis is a unique model for studying MMP function during vertebrate development because of its dependence on thyroid hormone (T3) and the ability to easily manipulate this process with exogenous T3. The MMP stromelysin-3 (ST3) is induced by T3, and its expression correlates with cell death during metamorphosis. We have previously shown that ST3 is both necessary and sufficient for larval epithelial cell death in the remodeling intestine. To investigate the roles of ST3 in other organs and especially on different cell types, we have analyzed the effect of transgenic overexpression of ST3 in the tail of premetamorphic tadpoles. We report for the first time that ST3 expression, in the absence of T3, caused significant muscle cell death in the tail of premetamorphic transgenic tadpoles. On the other hand, only relatively low levels of epidermal cell death were induced by precocious ST3 expression in the tail, contrasting what takes place during natural and T3-induced metamorphosis when ST3 expression is high. This cell type-specific apoptotic response to ST3 in the tail suggests distinct mechanisms regulating cell death in different tissues. Furthermore, our analyses of laminin receptor, an in vivo substrate of ST3 in the intestine, suggest that laminin receptor cleavage may be an underlying mechanism for the cell type-specific effects of ST3.The extracellular matrix (ECM),3 the dynamic milieu of the cell microenvironment, plays a critical role in dictating the fate of the cell. The cross-talk between the cell and ECM and the timely catabolism of the ECM are crucial for tissue remodeling during development (1). Matrix metalloproteinases (MMPs), extrinsic proteolytic regulators of the ECM, mediate this process to a large extent. MMPs are a large family of Zn2+-dependent endopeptidases potentially capable of cleaving the extracellular as well as nonextracellular proteins (29). The MMP superfamily includes collagenases, gelatinases, stromelysins, and membrane-type MMPs based on substrate specificity and domain organization (24). MMPs have been implicated to influence a wide range of physiological and pathological processes (1013). The roles of MMPs appear to be very complex. For example, MMPs have been suggested to play roles in both tumor promotion and suppression (1319). Unfortunately, relatively few functional studies have been carried out in vivo, especially in relation to the mechanisms involved during vertebrate development.Amphibian metamorphosis presents a fascinating experimental model to study MMP function during postembryonic development. A unique and salient feature of the metamorphic process is the absolute dependence on the signaling of thyroid hormone (2023). This makes it possible to prevent metamorphosis by simply inhibiting the synthesis of endogenous T3 or to induce precocious metamorphosis by merely adding physiological levels of T3 in the rearing water of premetamorphic tadpoles. Gene expression screens have identified the MMP stromelysin-3 (ST3) as a direct T3 response gene (2427). Expression studies have revealed a distinct spatial and temporal ST3 expression profile in correlation with metamorphic event, especially cell death (25, 2831). Organ culture studies on intestinal remodeling have directly substantiated an essential role of ST3 in larval epithelial cell death and ECM remodeling (32). Furthermore, precocious expression of ST3 alone in premetamorphic tadpoles through transgenesis is sufficient to induce ECM remodeling and larval epithelial apoptosis in the tadpole intestine (33). Thus, ST3 appears to be necessary and sufficient for intestinal epithelial cell death during metamorphosis.ST3 was first isolated as a breast cancer-associated gene (34), and unlike most other MMPs, ST3 is secreted as an active protease through a furin-dependent intracellular activation mechanism (35). Like many other MMPs, ST3 is expressed in a number of pathological processes, including most human carcinomas (11, 3640), as well as in many developmental processes in mammals (10, 34, 4143), although the physiological and pathological roles of ST3 in vivo are largely unknown in mammals. Interestingly, compared with other MMPs, ST3 has only weak activities toward ECM proteins in vitro but stronger activities against non-ECM proteins like α1 proteinase inhibitor and IGFBP-1 (4446). Although ST3 may cleave ECM proteins strongly in the in vivo environment, these findings suggest that the cleavage of non-ECM proteins is likely important for its biological roles. Consistently, we have recently identified a cell surface receptor, laminin receptor (LR) as an in vivo substrate of ST3 in the tadpole intestine during metamorphosis (4749). Analyses of LR expression and cleavage suggest that LR cleavage by ST3 is likely an important mechanism by which ST3 regulates the interaction between the larval epithelial cells and the ECM to induce cell death during intestinal remodeling (47, 48).Here, to investigate the role of ST3 in the apoptosis in other tissues during metamorphosis and whether LR cleavage serves as a mechanism for ST3 to regulate the fate of different cell types, we have analyzed the effects of precocious expression of ST3 in premetamorphic tadpole tail. The tail offers an opportunity to examine the effects of ST3 on different cell types. The epidermis, the fast and slow muscles, and the connective tissue underlying the epidermis in the myotendinous junctions and surrounding the notochord constitute the major tissue types in tail (50). Even though death is the destiny of all these cell types, it is not clear whether they all die through similar or different mechanisms. Microscopic and histochemical analyses have shown that at least the muscle and epidermal cells undergo T3-dependent apoptosis during metamorphosis (23, 29, 51, 52). To study whether ST3 regulates apoptosis of these two cell types, we have made use of the transgenic animals that express a transgenic ST3 under the control of a heat shock-inducible promoter (33). We show that whereas extensive apoptosis is present in both the epidermis and muscles during natural as well as T3-induced metamorphosis, transgenic expression of ST3 induces cell death predominantly in the muscles. Furthermore, we show that LR is expressed in the epidermis and connective tissue but not in muscles of the tadpole tail. More importantly, LR cleavage products are present in the tail during natural metamorphosis but not in transgenic tadpoles overexpressing ST3. These results suggest that ST3 has distinct effects on the epidermis and muscles in the tail, possibly because of the tissue-specific expression and function of LR.  相似文献   

17.
18.
Complex I (EC 1.6.99.3) of the bacterium Escherichia coli is considered to be the minimal form of the type I NADH dehydrogenase, the first enzyme complex in the respiratory chain. Because of its small size and relative simplicity, the E. coli enzyme has become a model used to identify and characterize the mechanism(s) by which cells regulate the synthesis and assembly of this large respiratory complex. To begin dissecting the processes by which E. coli cells regulate the expression of nuo and the assembly of complex I, we undertook a genetic analysis of the nuo locus, which encodes the 14 Nuo subunits comprising E. coli complex I. Here we present the results of studies, performed on an isogenic collection of nuo mutants, that focus on the physiological, biochemical, and molecular consequences caused by the lack of or defects in several Nuo subunits. In particular, we present evidence that NuoG, a peripheral subunit, is essential for complex I function and that it plays a role in the regulation of nuo expression and/or the assembly of complex I.

Complex I (NADH:ubiquinone oxidoreductase; EC 1.6.99.3), a type I NADH dehydrogenase that couples the oxidation of NADH to the generation of a proton motive force, is the first enzyme complex of the respiratory chain (2, 35, 47). The Escherichia coli enzyme, considered to be the minimal form of complex I, consists of 14 subunits instead of the 40 to 50 subunits associated with the homologous eukaryotic mitochondrial enzyme (17, 29, 30, 4850). E. coli also possesses a second NADH dehydrogenase, NDH-II, which does not generate a proton motive force (31). E. coli complex I resembles eukaryotic complex I in many ways (16, 17, 30, 49): it performs the same enzymatic reaction and is sensitive to a number of the same inhibitors, it consists of subunits homologous to those found in all proton-translocating NADH:ubiquinone oxidoreductases studied thus far, it is comprised of a large number of subunits relative to the number that comprise other respiratory enzymes, and it contains flavin mononucleotide and FeS center prosthetic groups. Additionally, it possesses an L-shaped topology (14, 22) like that of its Neurospora crassa homolog (27), and it consists of distinct fragments or subcomplexes. Whereas eukaryotic complex I can be dissected into a peripheral arm and a membrane arm, the E. coli enzyme consists of three subcomplexes referred to as the peripheral, connecting, and membrane fragments (29) (Fig. (Fig.1A).1A). The subunit composition of these three fragments correlates approximately with the organization of the 14 structural genes (nuoA to nuoN) (49) of the nuo (for NADH:ubiquinone oxidoreductase) locus (Fig. (Fig.1B),1B), an organization that is conserved in several other bacteria, including Salmonella typhimurium (3), Paracoccus denitrificans (53), Rhodobacter capsulatus (12), and Thermus thermophilus (54). The 5′ half of the locus contains a promoter (nuoP), previously identified and located upstream of nuoA (8, 49), and the majority of genes that encode subunits homologous to the nucleus-encoded subunits of eukaryotic complex I and to subunits of the Alcaligenes eutrophus NAD-reducing hydrogenase (17, 29, 30, 49). In contrast, the 3′ half contains the majority of the genes that encode subunits homologous to the mitochondrion-encoded subunits of eukaryotic complex I and to subunits of the E. coli formate-hydrogen lyase complex (17, 29, 30, 49). Whereas the nuclear homologs NuoE, NuoF, and NuoG constitute the peripheral fragment (also referred to as the NADH dehydrogenase fragment [NDF]), the nuclear homologs NuoB, NuoC, NuoD, and NuoI constitute the connecting fragment. The mitochondrial homologs NuoA, NuoH, NuoJ, NuoK, NuoL, NuoM, and NuoN constitute the membrane fragment (29). E. coli complex I likely evolved by fusion of preexisting protein assemblies constituting modules for electron transfer and proton translocation (1719, 30). Open in a separate windowFIG. 1Schematic of E. coli complex I and the nuo locus. Adapted with permission of the publisher (17, 29, 30, 49). (A) E. coli complex I is comprised of three distinct fragments: the peripheral (light gray), connecting (white), and membrane (dark gray) fragments (17, 29). The peripheral fragment (NDF) is comprised of the nuclear homologs NuoE, -F, and -G and exhibits NADH dehydrogenase activity that oxidizes NADH to NAD+; the connecting fragment is comprised of the nuclear homologs NuoB, -C, -D, and -I; and the membrane fragment is comprised of the mitochondrial homologs NuoA, -H, and -J to -N and catalyzes ubiquinone (Q) to its reduced form (QH2). FMN, flavin mononucleotide. (B) The E. coli nuo locus encodes the 14 Nuo subunits that constitute complex I. The 5′ half of the locus contains a previously identified promoter (nuoP) and the majority of genes that encode the peripheral and connecting subunits (light gray and white, respectively). The 3′ half of the locus contains the majority of the genes encoding the membrane subunits (dark gray). The 3′ end of nuoG encodes a C-Terminal region (CTR) of the NuoG subunit (hatched).Because of its smaller size and relative simplicity, researchers recently have begun to utilize complex I of E. coli, and that of its close relative S. typhimurium, to identify and characterize the mechanism(s) by which cells regulate the synthesis and assembly of this large respiratory complex (3, 8, 46) and to investigate the diverse physiological consequences caused by defects in this enzyme (4, 6, 10, 40, 59). Such defects affect the ability of cells to perform chemotaxis (40), to grow on certain carbon sources (4, 6, 10, 40, 57), to survive stationary phase (59), to perform energy-dependent proteolysis (4), to regulate the expression of at least one gene (32), and to maintain virulence (5).To begin dissecting the processes by which E. coli cells regulate the expression of nuo and the assembly of complex I, we undertook a genetic analysis of the nuo locus. Here, we present the results of studies, performed on an isogenic collection of nuo mutants, that focus on the physiological, biochemical, and molecular consequences caused by the lack of or defects in several Nuo subunits. In particular, we present evidence that NuoG, a peripheral subunit, is essential for complex I function and that it plays a role in the regulation of nuo expression and/or the assembly of complex I.  相似文献   

19.
20.
A polyomavirus mutant (315YF) blocked in binding phosphatidylinositol 3-kinase (PI 3-kinase) has previously been shown to be partially deficient in transformation and to induce fewer tumors and with a significant delay compared to wild-type virus. The role of polyomavirus middle T antigen-activated PI 3-kinase in apoptosis was investigated as a possible cause of this behavior. When grown in medium containing 1d-3-deoxy-3-fluoro-myo-inositol to block formation of 3′-phosphorylated phosphatidylinositols, F111 rat fibroblasts transformed by wild-type polyomavirus (PyF), but not normal F111 cells, showed a marked loss of viability with evidence of apoptosis. Similarly, treatment with wortmannin, an inhibitor of PI 3-kinase, stimulated apoptosis in PyF cells but not in normal cells. Activation of Akt, a serine/threonine kinase whose activity has been correlated with regulation of apoptosis, was roughly twofold higher in F111 cells transformed by either wild-type virus or mutant 250YS blocked in binding Shc compared to cells transformed by mutant 315YF. In the same cells, levels of apoptosis were inversely correlated with Akt activity. Apoptosis induced by serum withdrawal in Rat-1 cells expressing a temperature-sensitive p53 was shown to be at least partially p53 independent. Expression of either wild-type or 250YS middle T antigen inhibited apoptosis in serum-starved Rat-1 cells at both permissive and restrictive temperatures for p53. Mutant 315YF middle T antigen was partially defective for inhibition of apoptosis in these cells. The results indicate that unlike other DNA tumor viruses which block apoptosis by inactivation of p53, polyomavirus achieves protection from apoptotic death through a middle T antigen–PI 3-kinase–Akt pathway that is at least partially p53 independent.Programmed cell death occurs during normal development and under certain pathological conditions. In mammalian cells, apoptosis can be induced by a variety of stimuli, including DNA damage (45), virus infection (54, 57), oncogene activation (25), and serum withdrawal (34, 37). Apoptosis can also be blocked by a number of factors, including adenovirus E1B 55- or 19-kDa proteins (9, 16), baculovirus p35 and iap genes (10), Bcl-2 (36, 61), and survival factors (12, 21). DNA tumor viruses have evolved mechanisms that both trigger and inhibit apoptosis. These frequently involve binding and inactivation of tumor suppressor proteins. E7 in some papillomaviruses (22), E1A in adenovirus (31, 43, 64), and large T antigen in simian virus 40 (SV40) (17) bind Rb and/or p300 and lead to upregulation of p53, which is thought to trigger apoptosis in virus-infected cells. The same viruses also inhibit apoptosis by inactivating p53 by various mechanisms (44, 63, 67). In contrast, the mechanism by which polyomavirus interacts with apoptotic pathways in the cell is not known; no direct interaction with p53 by any of the proteins encoded by this virus has been demonstrated (19, 62).The principal oncoprotein of polyomavirus is the middle T antigen. Neoplastic transformation by polyomavirus middle T antigen has as a central feature its association with and activation of members of the Src family of tyrosine kinases p60c-src (13) and p62c-yes (42). The major known consequence of these interactions is phosphorylation of middle T antigen on specific tyrosine residues creating binding sites for other signaling proteins. Phosphorylation at tyrosines 250, 315, and 322 promotes binding to Shc (18), the p85 regulatory subunit of phosphatidylinositol 3-kinase (PI 3-kinase) (59), and phospholipase Cγ-1 (58), respectively. Recognition of multiple signaling pathways emanating from middle T antigen has led to a keen interest in identifying their downstream biochemical effects, which collectively lead to the emergence of neoplastic transformation and presumably underlie the dramatic ability of the virus to induce many kinds of tumors in the mouse.Previous work has shown that the binding of PI 3-kinase to middle T antigen is essential for full transformation of rat fibroblasts in culture (8) and for rapid development of a broad spectrum of tumors in mice (30), for translocation of the GLUT1 transporter (68), and activation of p70 S6 kinase (14). While the mutant 315YF (blocked in PI 3-kinase activation) was able to induce some tumors, it did so at reduced frequencies and with an average latency three times longer than that of either the wild-type virus or a mutant, 250YS, blocked in binding Shc (4, 30). Recent studies have indicated a role of PI 3-kinase in blocking apoptosis in nonviral systems. Growth factor receptors acting through protein tyrosine kinases may prevent apoptosis by activating PI 3-kinase in PC12 cells, T lymphocytes, hematopoietic progenitors, and rat fibroblasts (7, 48, 56, 65, 66). The failure of mutant 315YF to induce full transformation of cells in culture and to induce the rapid development of tumors in mice could therefore be related, at least in part, to a failure to block apoptosis. In this study, we focus on the question of whether middle T antigen–PI 3-kinase interaction is involved in blocking apoptosis in cells transformed by polyomavirus.  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