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.
2.
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   

3.
The exponential growth in the volume of publications in the biomedical domain has made it impossible for an individual to keep pace with the advances. Even though evidence-based medicine has gained wide acceptance, the physicians are unable to access the relevant information in the required time, leaving most of the questions unanswered. This accentuates the need for fast and accurate biomedical question answering systems. In this paper we introduce INDOC—a biomedical question answering system based on novel ideas of indexing and extracting the answer to the questions posed. INDOC displays the results in clusters to help the user arrive the most relevant set of documents quickly. Evaluation was done against the standard OHSUMED test collection. Our system achieves high accuracy and minimizes user effort.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24]  相似文献   

4.
5.
6.
Early onset generalized dystonia (DYT1) is an autosomal dominant neurological disorder caused by deletion of a single glutamate residue (torsinA ΔE) in the C-terminal region of the AAA+ (ATPases associated with a variety of cellular activities) protein torsinA. The pathogenic mechanism by which torsinA ΔE mutation leads to dystonia remains unknown. Here we report the identification and characterization of a 628-amino acid novel protein, printor, that interacts with torsinA. Printor co-distributes with torsinA in multiple brain regions and co-localizes with torsinA in the endoplasmic reticulum. Interestingly, printor selectively binds to the ATP-free form but not to the ATP-bound form of torsinA, supporting a role for printor as a cofactor rather than a substrate of torsinA. The interaction of printor with torsinA is completely abolished by the dystonia-associated torsinA ΔE mutation. Our findings suggest that printor is a new component of the DYT1 pathogenic pathway and provide a potential molecular target for therapeutic intervention in dystonia.Early onset generalized torsion dystonia (DYT1) is the most common and severe form of hereditary dystonia, a movement disorder characterized by involuntary movements and sustained muscle spasms (1). This autosomal dominant disease has childhood onset and its dystonic symptoms are thought to result from neuronal dysfunction rather than neurodegeneration (2, 3). Most DYT1 cases are caused by deletion of a single glutamate residue at positions 302 or 303 (torsinA ΔE) of the 332-amino acid protein torsinA (4). In addition, a different torsinA mutation that deletes amino acids Phe323–Tyr328 (torsinA Δ323–328) was identified in a single family with dystonia (5), although the pathogenic significance of this torsinA mutation is unclear because these patients contain a concomitant mutation in another dystonia-related protein, ϵ-sarcoglycan (6). Recently, genetic association studies have implicated polymorphisms in the torsinA gene as a genetic risk factor in the development of adult-onset idiopathic dystonia (7, 8).TorsinA contains an N-terminal endoplasmic reticulum (ER)3 signal sequence and a 20-amino acid hydrophobic region followed by a conserved AAA+ (ATPases associated with a variety of cellular activities) domain (9, 10). Because members of the AAA+ family are known to facilitate conformational changes in target proteins (11, 12), it has been proposed that torsinA may function as a molecular chaperone (13, 14). TorsinA is widely expressed in brain and multiple other tissues (15) and is primarily associated with the ER and nuclear envelope (NE) compartments in cells (1620). TorsinA is believed to mainly reside in the lumen of the ER and NE (1719) and has been shown to bind lamina-associated polypeptide 1 (LAP1) (21), lumenal domain-like LAP1 (LULL1) (21), and nesprins (22). In addition, recent evidence indicates that a significant pool of torsinA exhibits a topology in which the AAA+ domain faces the cytoplasm (20). In support of this topology, torsinA is found in the cytoplasm, neuronal processes, and synaptic terminals (2, 3, 15, 2326) and has been shown to bind cytosolic proteins snapin (27) and kinesin light chain 1 (20). TorsinA has been proposed to play a role in several cellular processes, including dopaminergic neurotransmission (2831), NE organization and dynamics (17, 22, 32), and protein trafficking (27, 33). However, the precise biological function of torsinA and its regulation remain unknown.To gain insights into torsinA function, we performed yeast two-hybrid screens to search for torsinA-interacting proteins in the brain. We report here the isolation and characterization of a novel protein named printor (protein interactor of torsinA) that interacts selectively with wild-type (WT) torsinA but not the dystonia-associated torsinA ΔE mutant. Our data suggest that printor may serve as a cofactor of torsinA and provide a new molecular target for understanding and treating dystonia.  相似文献   

7.
8.
9.
10.
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   

11.
Group I metabotropic glutamate receptors (mGluRs) are coupled via phospholipase Cβ to the hydrolysis of phosphoinositides and function to modulate neuronal excitability and synaptic transmission at glutamatergic synapses. The desensitization of Group I mGluR signaling is thought to be mediated primarily via second messenger-dependent protein kinases and G protein-coupled receptor kinases. We show here that both mGluR1 and mGluR5 interact with the calcineurin inhibitor protein (CAIN). CAIN is co-immunoprecipitated in a complex with Group I mGluRs from both HEK 293 cells and mouse cortical brain lysates. Purified CAIN and its C-terminal domain specifically interact with glutathione S-transferase fusion proteins corresponding to the second intracellular loop and the distal C-terminal tail domains of mGluR1. The interaction of CAIN with mGluR1 could also be blocked using a Tat-tagged peptide corresponding to the mGluR1 second intracellular loop domain. Overexpression of full-length CAIN attenuates the agonist-stimulated endocytosis of both mGluR1a and mGluR5a in HEK 293 cells, but expression of the CAIN C-terminal domain does not alter mGluR5a internalization. In contrast, overexpression of either full-length CAIN or the CAIN C-terminal domain impairs agonist-stimulated inositol phosphate formation in HEK 293 cells expressing mGluR1a. This CAIN-mediated antagonism of mGluR1a signaling appears to involve the disruption of receptor-Gαq/11 complexes. Taken together, these observations suggest that the association of CAIN with intracellular domains involved in mGluR/G protein coupling provides an additional mechanism by which Group I mGluR endocytosis and signaling are regulated.Metabotropic glutamate receptors (mGluRs)2 play an essential role in regulating neuronal plasticity, development, and neurotoxicity and belong to the G protein-coupled receptor superfamily of integral membrane proteins (14). The mGluR family can be subclassified into three groups based on sequence homology, G protein specificity, and pharmacology. Group I mGluRs (mGluR1 and mGluR5) couple via the heterotrimeric Gαq/11 proteins to the activation of phospholipase Cβ, resulting in the formation of inositol 1,4,5-triphosphate and diacylglycerol, the release of Ca2+ from intracellular stores, and the activation of protein kinase C (PKC) (46).The regulation of mGluR signal transduction involves numerous proteins that function to regulate signaling at both the level of the heterotrimeric G protein and the receptor (68). At the level of the receptor, Group I mGluR activity is regulated by a process termed desensitization, which protects against both acute and chronic receptor overstimulation (9, 10). The attenuation of Group I mGluR signaling can be mediated by both phosphorylation-dependent and phosphorylation-independent processes (11). The phosphorylation-independent attenuation of Group I mGluR signaling is mediated by GRK2 (G protein-coupled receptor kinase 2), which is composed of three functional domains: an N-terminal RGS (regulator of G protein signaling) homology domain, a central catalytic domain, and a C-terminal Gβγ-binding pleckstrin homology domain (12). GRK2-mediated desensitization of Group I mGluRs does not require catalytic activity but rather requires the interaction of the GRK2 RGS homology domain with both the second intracellular loop domain of mGluR1 and the α-subunit of Gαq/11, thereby attenuating heterotrimeric G protein coupling (1315). Phosphorylation-independent desensitization of mGluR1 signaling is also mediated by optineurin, an effect that is enhanced by the expression of mutant huntingtin (16). Phosphorylation-dependent desensitization of Group I mGluR responsiveness involves the phosphorylation of PKC consensus sequence localized within the intracellular loop and C-terminal tail domains of mGluR1 and mGluR5 by PKC (17, 18). It is proposed that calcineurin and mGluR5 may exist in a signaling complex in the brain and that calcineurin may function to modulate mGluR5 signaling by directly dephosphorylating the receptor at a PKC consensus site that contributes to mGluR5 desensitization (19). Calcineurin is also linked to the regulation of endocytosis via its interaction with dynamin-1 (20).On the basis of the observation that calcineurin may form a complex with Group I mGluRs, we hypothesized that CAIN (calcineurin inhibitor protein) might also interact with Group I mGluRs and modulate their endocytosis and signaling. CAIN, also known as Cabin1 (calcineurin-binding protein), was first identified as a protein that binds to calcineurin and was shown to inhibit calcineurin catalytic activity (2123). Previous studies also demonstrated that CAIN may interact with amphiphysin-1, dynamin-1, and α-adaptin and led to the suggestion that CAIN functions as a component of synaptic endocytic complexes (24). Consistent with this hypothesis, the overexpression of CAIN in human embryonic kidney (HEK 293) cells resulted in attenuated transferrin receptor endocytosis.We show here that CAIN interacts with the second intracellular loop and C-terminal tail domains of Group I mGluRs, inhibits Group I mGluR internalization, and attenuates mGluR1a signaling by disrupting receptor-Gαq/11 complexes. Taken together, these results describe an additional mechanism by which Group I mGluR activity may be regulated.  相似文献   

12.
A variety of high-throughput methods have made it possible to generate detailed temporal expression data for a single gene or large numbers of genes. Common methods for analysis of these large data sets can be problematic. One challenge is the comparison of temporal expression data obtained from different growth conditions where the patterns of expression may be shifted in time. We propose the use of wavelet analysis to transform the data obtained under different growth conditions to permit comparison of expression patterns from experiments that have time shifts or delays. We demonstrate this approach using detailed temporal data for a single bacterial gene obtained under 72 different growth conditions. This general strategy can be applied in the analysis of data sets of thousands of genes under different conditions.[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   

13.
The inference of gene regulatory networks is a key issue for genomic signal processing. This paper addresses the inference of probabilistic Boolean networks (PBNs) from observed temporal sequences of network states. Since a PBN is composed of a finite number of Boolean networks, a basic observation is that the characteristics of a single Boolean network without perturbation may be determined by its pairwise transitions. Because the network function is fixed and there are no perturbations, a given state will always be followed by a unique state at the succeeding time point. Thus, a transition counting matrix compiled over a data sequence will be sparse and contain only one entry per line. If the network also has perturbations, with small perturbation probability, then the transition counting matrix would have some insignificant nonzero entries replacing some (or all) of the zeros. If a data sequence is sufficiently long to adequately populate the matrix, then determination of the functions and inputs underlying the model is straightforward. The difficulty comes when the transition counting matrix consists of data derived from more than one Boolean network. We address the PBN inference procedure in several steps: (1) separate the data sequence into "pure" subsequences corresponding to constituent Boolean networks; (2) given a subsequence, infer a Boolean network; and (3) infer the probabilities of perturbation, the probability of there being a switch between constituent Boolean networks, and the selection probabilities governing which network is to be selected given a switch. Capturing the full dynamic behavior of probabilistic Boolean networks, be they binary or multivalued, will require the use of temporal data, and a great deal of it. This should not be surprising given the complexity of the model and the number of parameters, both transitional and static, that must be estimated. In addition to providing an inference algorithm, this paper demonstrates that the data requirement is much smaller if one does not wish to infer the switching, perturbation, and selection probabilities, and that constituent-network connectivity can be discovered with decent accuracy for relatively small time-course sequences.[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]  相似文献   

14.
15.
16.
17.
Distribution-free statistical tests offer clear advantages in situations where the exact unadjusted -values are required as input for multiple testing procedures. Such situations prevail when testing for differential expression of genes in microarray studies. The Cramér-von Mises two-sample test, based on a certain -distance between two empirical distribution functions, is a distribution-free test that has proven itself as a good choice. A numerical algorithm is available for computing quantiles of the sampling distribution of the Cramér-von Mises test statistic in finite samples. However, the computation is very time- and space-consuming. An counterpart of the Cramér-von Mises test represents an appealing alternative. In this work, we present an efficient algorithm for computing exact quantiles of the -distance test statistic. The performance and power of the -distance test are compared with those of the Cramér-von Mises and two other classical tests, using both simulated data and a large set of microarray data on childhood leukemia. The -distance test appears to be nearly as powerful as its counterpart. The lower computational intensity of the -distance test allows computation of exact quantiles of the null distribution for larger sample sizes than is possible for the Cramér-von Mises test.[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]  相似文献   

18.
19.
A trehalose synthase (TSase) that catalyzes the synthesis of trehalose from d-glucose and α-d-glucose 1-phosphate (α-d-glucose 1-P) was detected in a basidiomycete, Grifola frondosa. TSase was purified 106-fold to homogeneity with 36% recovery by ammonium sulfate precipitation and several steps of column chromatography. The native enzyme appears to be a dimer since it has apparent molecular masses of 120 kDa, as determined by gel filtration column chromatography, and 60 kDa, as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Although TSase catalyzed the phosphorolysis of trehalose to d-glucose and α-d-glucose 1-P, in addition to the synthesis of trehalose from the two substrates, the TSase equilibrium strongly favors trehalose synthesis. The optimum temperatures for phosphorolysis and synthesis of trehalose were 32.5 to 35°C and 35 to 37.5°C, respectively. The optimum pHs for these reactions were 6.5 and 6.5 to 6.8, respectively. The substrate specificity of TSase was very strict: among eight disaccharides examined, only trehalose was phosphorolyzed, and only α-d-glucose 1-P served as a donor substrate with d-glucose as the acceptor in trehalose synthesis. Two efficient enzymatic systems for the synthesis of trehalose from sucrose were identified. In system I, the α-d-glucose 1-P liberated by 1.05 U of sucrose phosphorylase was linked with d-glucose by 1.05 U of TSase, generating trehalose at the initial synthesis rate of 18 mmol/h in a final yield of 90 mol% under optimum conditions (300 mM each sucrose and glucose, 20 mM inorganic phosphate, 37.5°C, and pH 6.5). In system II, we added 1.05 U of glucose isomerase and 20 mM MgSO4 to the reaction mixture of system I to convert fructose, a by-product of the sucrose phosphorylase reaction, into glucose. This system generated trehalose at the synthesis rate of 4.5 mmol/h in the same final yield.Trehalose (1-α-d-glucopyranosyl-α-d-glucopyranoside) is a nonreducing disaccharide with an α,α-1,1 glycosidic linkage and is widely distributed in plants, insects, fungi, yeast, and bacteria (7). Due to the absence of reducing ends in trehalose, it is highly resistant to heat, pH, and Maillard’s reaction (24). In trehalose-producing organisms, this compound may serve as an energy reserve, a buffer against stresses such as desiccation and freezing, and a protein stabilizer (5, 7, 26, 31, 32). If trehalose can be produced economically, then it has potential commercial applications as a sweetener, a food stabilizer, and an additive in cosmetics and pharmaceuticals (6, 25). Recently, trehalose production through fermentation of yeast (17) and Corynebacterium (30), enzymatic processes from starch (18, 34) and maltose (19, 22, 23, 33), and extraction from transformed plants (10) has been reported.Our approach to trehalose production is to use an enzymatic process to produce trehalose from sucrose, one of the least expensive sugars. Since sucrose is efficiently converted to α-d-glucose 1-phosphate (α-d-glucose 1-P) and fructose by sucrose phosphorylase (SPase), we screened microorganisms for an enzyme that converts α-d-glucose 1-P to trehalose on the assumption that the combination of the putative trehalose synthase (TSase) and SPase would convert sucrose into trehalose. Although similar enzyme activities have been reported in the basidiomycete Flammulina velutipes (11) and in the yeast Pichia fermentans (27), these enzymes have not been well characterized.Our objectives were (i) to screen microorganisms, primarily fungi, for TSase activity; (ii) to purify and characterize the TSase; (iii) to identify the enzymatic process by which trehalose is produced from sucrose; and (iv) to identify an enzymatic process for production of trehalose from sucrose in which the fructose component is also converted to trehalose.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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