首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Therapeutic agents are urgently required to cure several common and fatal neurodegenerative disorders caused by protein misfolding and aggregation, including amyotrophic lateral sclerosis (ALS), Parkinson’s disease (PD), and Alzheimer’s disease (AD). Protein disaggregases that reverse protein misfolding and restore proteins to native structure, function, and localization could mitigate neurodegeneration by simultaneously reversing 1) any toxic gain of function of the misfolded form and 2) any loss of function due to misfolding. Potentiated variants of Hsp104, a hexameric AAA+ ATPase and protein disaggregase from yeast, have been engineered to robustly disaggregate misfolded proteins connected with ALS (e.g., TDP-43 and FUS) and PD (e.g., α-synuclein). However, Hsp104 has no metazoan homologue. Metazoa possess protein disaggregase systems distinct from Hsp104, including Hsp110, Hsp70, and Hsp40, as well as HtrA1, which might be harnessed to reverse deleterious protein misfolding. Nevertheless, vicissitudes of aging, environment, or genetics conspire to negate these disaggregase systems in neurodegenerative disease. Thus, engineering potentiated human protein disaggregases or isolating small-molecule enhancers of their activity could yield transformative therapeutics for ALS, PD, and AD.We urgently need to pioneer game-changing solutions to remedy a number of increasingly prevalent and fatal neurodegenerative disorders, including amyotrophic lateral sclerosis (ALS), Parkinson’s disease (PD), and Alzheimer’s disease (AD; Cushman et al., 2010 ; Jackrel and Shorter, 2015 ). These disorders relentlessly erode our morale and economic resources. Aging is the major risk factor for all of these diseases, which threaten public health on a global scale and represent a severe impediment to living longer lives. A number of promising drugs have emerged to treat cancer and heart disease, but, distressingly, this is not the case for these and other neurodegenerative diseases, for which drug pipelines lie dormant and empty. This situation is unacceptable, and an impending healthcare crisis looms worldwide as population demographics inexorably shift toward older age groups.ALS, PD, AD, and related neurodegenerative disorders are unified by a common underlying theme: the misfolding and aggregation of specific proteins (characteristic for each disease) in the CNS (Cushman et al., 2010 ; Eisele et al., 2015 ). Thus, in ALS, typically an RNA-binding protein with a prion-like domain, TDP-43, mislocalizes from the nucleus to cytoplasmic inclusions in degenerating motor neurons (Neumann et al., 2006 ; Gitler and Shorter, 2011 ; King et al., 2012 ; Robberecht and Philips, 2013 ; March et al., 2016 ). In PD, α-synuclein forms toxic soluble oligomers and cytoplasmic aggregates, termed Lewy bodies, in degenerating dopaminergic neurons (Dehay et al., 2015 ). By contrast, in AD, amyloid-β (Aβ) peptides form extracellular plaques and the microtubule-binding protein, tau, forms cytoplasmic neurofibrillary tangles in afflicted brain regions (Jucker and Walker, 2011 ). Typically, these disorders are categorized into ∼80–90% sporadic cases and ∼10–20% familial cases. Familial forms of disease often have clear genetic causes, which might one day be amenable to gene editing via clustered regularly interspaced short palindromic repeats (CRISPR)-Cas9 therapeutics if critical safety and ethical concerns can be successfully addressed and respected (Doudna and Charpentier, 2014 ; Baltimore et al., 2015 ; Rahdar et al., 2015 ; Callaway, 2016 ). However, the more common sporadic forms of disease often have no clear underlying genetics, and wild-type proteins misfold (Cushman et al., 2010 ; Jucker and Walker, 2011 ; Robberecht and Philips, 2013 ; Dehay et al., 2015 ). Consequently, therapeutic agents that directly target and safely reverse deleterious protein misfolding are likely to have broad utility (Eisele et al., 2015 ).There are no treatments that directly target the reversal of the protein-misfolding phenomena that underlie these disorders (Jackrel and Shorter, 2015 ). Strategies that directly reverse protein misfolding and restore proteins to native form and function could, in principle, eradicate any severely damaging loss-of-function or toxic gain-of-function phenotypes caused by misfolded conformers (Figure 1; Jackrel and Shorter, 2015 ). Moreover, therapeutic disaggregases would dismantle self-templating amyloid or prion structures, which spread pathology and nucleate formation of neurotoxic, soluble oligomers (Figure 1; Cushman et al., 2010 ; Cohen et al., 2013 ; Guo and Lee, 2014 ; Jackrel and Shorter, 2015 ). My group has endeavored to engineer and evolve Hsp104, a hexameric AAA+ ATPase and protein disaggregase from yeast (DeSantis and Shorter, 2012 ; Sweeny and Shorter, 2015 ), to more effectively disaggregate misfolded proteins connected with various neurodegenerative disorders, including ALS (e.g., TDP-43 and FUS) and PD (e.g., α-synuclein). Although wild-type Hsp104 can disaggregate diverse amyloid and prion conformers, as well as toxic soluble oligomers (Lo Bianco et al., 2008 ; DeSantis et al., 2012 ), its activity against human neurodegenerative disease proteins is suboptimal. Is it even possible to improve on existing Hsp104 disaggregase activity, which has been wrought over the course of millions of years of evolution?Open in a separate windowFIGURE 1:Therapeutic protein disaggregases. Two malicious problems are commonly associated with protein misfolding into disordered aggregates, toxic oligomers, and cross–β amyloid or prion fibrils: 1) a toxic gain of function of the protein in various misfolded states; and 2) a loss of function of the protein in the various misfolded states. These problems can contribute to the etiology of diverse neurodegenerative diseases in a combinatorial or mutually exclusive manner. A therapeutic protein disaggregase would reverse protein misfolding and recover natively folded functional proteins from disordered aggregates, toxic oligomers, and cross–β amyloid or prion fibrils. In this way, any toxic gain of function or toxic loss of function caused by protein misfolding would be simultaneously reversed. Ideally, all toxic misfolded conformers would be purged. Therapeutic protein disaggregases could thus have broad utility for various fatal neurodegenerative diseases.Remarkably, the answer to this question is yes! We used nimble yeast models of neurodegenerative proteinopathies (Outeiro and Lindquist, 2003 ; Gitler, 2008 ; Johnson et al., 2008 ; Sun et al., 2011 ; Khurana et al., 2015 ) as a platform to isolate enhanced disaggregases from large libraries of Hsp104 variants generated by error-prone PCR (Jackrel et al., 2014b ). In this way, we reprogrammed Hsp104 to yield the first disaggregases that reverse TDP-43, FUS (another RNA-binding protein with a prion-like domain connected to ALS), and α-synuclein (connected to PD) aggregation and proteotoxicity (Jackrel et al., 2014a , 2015 ; Jackrel and Shorter, 2015 ; Torrente et al., 2016 ). Remarkably, a therapeutic gain of Hsp104 function could be elicited by a single missense mutation (Jackrel et al., 2014a , 2015 ; Jackrel and Shorter, 2015 ). Under conditions in which Hsp104 was ineffective, potentiated Hsp104 variants dissolved protein inclusions, restored protein localization (e.g., TDP-43 returned to the nucleus from cytoplasmic inclusions), suppressed proteotoxicity, and attenuated dopaminergic neurodegeneration in a Caenorhabditis elegans PD model (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ). Remarkably, these therapeutic modalities originated from degenerate loss of amino acid identity at select positions of Hsp104 rather than specific mutation (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ). Some of these changes were remarkably small (Jackrel et al., 2014a , 2015 ; Jackrel and Shorter, 2015 ). Thus, potentiated Hsp104 variants could be generated by removal of a methyl group from a single side chain or addition or removal of a single methylene bridge from a single side chain (Jackrel et al., 2014a , 2015 ; Jackrel and Shorter, 2015 ). Thus, small molecules that bind in accessible regions of Hsp104 rich in potentiating mutations might also be able to enhance activity. However, a small-scale screen for small-molecule modulators of Hsp104 revealed only inhibitors (Torrente et al., 2014 ). Nonetheless, our work has established that disease-associated aggregates and amyloid are tractable targets and that enhanced artificial disaggregases can restore proteostasis and mitigate neurodegeneration (Jackrel and Shorter, 2015 ).One surprising aspect of this work is just how many Hsp104 variants we could isolate with potentiated activity. We now have hundreds (Jackrel et al., 2014a ; Jackrel et al., 2015 ). Typically, potentiated Hsp104 variants displayed enhanced activity against several neurodegenerative disease proteins. For example, Hsp104A503S rescued the aggregation and toxicity of TDP-43, FUS, TAF15, and α-synuclein (Jackrel et al., 2014a ; Jackrel and Shorter, 2014 ). By contrast, some potentiated Hsp104 variants rescued only the aggregation and toxicity of a subset of disease proteins. For example, Hsp104D498V rescued only the aggregation and toxicity of FUS and α-synuclein (Jackrel et al., 2014a ). A challenge that lies ahead is to engineer potentiated Hsp104 variants that are highly substrate specific to mitigate any potential off-target effects, should they arise (Jackrel and Shorter, 2015 ).Very small changes in primary sequence yield potentiated Hsp104 variants. However, Hsp104 has no metazoan homologue (Erives and Fassler, 2015 ). Now comes the important point. Neuroprotection could be broadly achieved by making very subtle modifications to existing human chaperones with newly appreciated disaggregase activity—for example, Hsp110, Hsp70, and Hsp40 (Torrente and Shorter, 2013 ) and HtrA1 (Poepsel et al., 2015 ).Whether Metazoa even possess a powerful protein disaggregation and reactivation machinery had been a long-standing enigma (Torrente and Shorter, 2013 ). However, it has recently emerged that two metazoan chaperone systems—1) Hsp110, Hsp70, and Hsp40 (Torrente and Shorter, 2013 ) and 2) HtrA1 (Poepsel et al., 2015 )—possess disaggregase activity that could be therapeutically harnessed or stimulated to reverse deleterious protein misfolding in neurodegenerative disease. I suspect that Metazoa harbor additional disaggregase systems that remain to be identified (Guo et al., 2014 ). Whether due to vicissitudes of aging, environment, or genetic background, these disaggregase systems fail in the context of ALS, PD, and AD. Based on the surprising precedent of our potentiated versions of Hsp104 (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ), I hypothesize that it is possible to engineer and evolve potentiated variants of these human protein disaggregases to more effectively counter deleterious misfolding events in ALS, PD, and AD (Torrente and Shorter, 2013 ; Mack and Shorter, 2016 ).Using classical biochemical reconstitution, it was discovered that one mammalian protein-disaggregase system comprises three molecular chaperones—Hsp110, Hsp70, and Hsp40—which synergize to dissolve and reactivate model proteins trapped in disordered aggregates and can even depolymerize amyloid fibrils formed by α-synuclein from their ends (Shorter, 2011 ; Duennwald et al., 2012 ; Torrente and Shorter, 2013 ). Hsp110, Hsp70, and Hsp40 isoforms are found in the cytoplasm, nucleus, and endoplasmic reticulum, which suggest that protein disaggregation and reactivation can occur in several compartments (Finka et al., 2015 ). Subsequent studies suggest that this system may be more powerful than initially anticipated (Rampelt et al., 2012 ; Mattoo et al., 2013 ; Gao et al., 2015 ; Nillegoda et al., 2015 ). Nonetheless, this system must become overwhelmed in neurodegenerative disorders. Perhaps selectively vulnerable neurons display particular deficits in this machinery. Hence, potentiating the activity of this system via engineering could be extremely valuable. It is promising that directed evolution studies yielded DnaK (Hsp70 in Escherichia coli) variants with improved ability to refold specific substrates (Aponte et al., 2010 ; Schweizer et al., 2011 ; Mack and Shorter, 2016 ), but whether this can be extended to human Hsp70 and the disaggregation of neurodegenerative disease proteins is uncertain.It is exciting that recent studies have revealed that HtrA1, a homo-oligomeric PDZ serine protease, can dissolve and degrade AD-linked tau and Aβ42 fibrils in an ATP-independent manner (Tennstaedt et al., 2012 ; Poepsel et al., 2015 ). HtrA1 first dissolves tau and Aβ42 fibrils and then degrades them, as protease-defective HtrA1 variants dissolve fibrils without degrading them (Poepsel et al., 2015 ). HtrA1 is found in the cytoplasm (∼30%) but is also secreted (∼70%; Poepsel et al., 2015 ). Indeed, HtrA1 is known to degrade substrates in both the extracellular space and the cytoplasm (Chien et al., 2009 ; Campioni et al., 2010 ; Tiaden and Richards, 2013 ). Thus HtrA1 could dissolve Aβ42 fibrils in the extracellular space and tau fibrils in the cytoplasm and simultaneously destroy the two cardinal features of AD (Poepsel et al., 2015 ). I suspect that this system becomes overwhelmed or is insufficient in AD, and thus potentiating and tailoring HtrA1 disaggregase activity could be a valuable therapeutic strategy. For example, it might be advantageous to simply degrade Aβ42 after disaggregation if the peptide has no beneficial function. Thus HtrA1 variants with enhanced disaggregation and degradation activity against Aβ42 could be extremely useful. However, Aβ42 (and related Aβ peptides) may have physiological functions that are presently underappreciated (Soscia et al., 2010 ; Fedele et al., 2015 ), in which case HtrA1 variants with enhanced disaggregase activity but reduced proteolytic activity could be vital. HtrA1 variants with enhanced disaggregase activity but reduced proteolytic activity may also be particularly important to recover functional tau from neurofibrillary tangles to reverse loss of tau function in AD and various tauopathies (Santacruz et al., 2005 ; Trojanowski and Lee, 2005 ; Dixit et al., 2008 ).I suggest that relatively small changes in primary sequence will yield large increases in disaggregase activity for these systems as they do for Hsp104 (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ). If true, this would further suggest that small molecules that bind in the appropriate regions of Hsp110, Hsp70, Hsp40, or HtrA1 might also enhance disaggregase activity. Thus, isolating small-molecule enhancers of the Hsp110, Hsp70, and Hsp40 or HtrA1 disaggregase systems could yield important therapeutics. Indeed, I hypothesize that enhancing the activity of the Hsp110, Hsp70, and Hsp40 or HtrA1 disaggregase system with specific small molecules will enable dissolution of toxic oligomeric and amyloid forms of various disease proteins and confer therapeutic benefits in ALS, PD, AD, and potentially other neurodegenerative disorders.It is intriguing that several small molecules are already known to enhance various aspects of Hsp70 chaperone activity (Pratt et al., 2015 ; Shrestha et al., 2016 ). These include MKT-077, JG-98, YM-1, YM-8, and 115-7c (Wisen et al., 2010 ; Pratt et al., 2015 ). It is not known whether any of these stimulates the disaggregase activity of the Hsp110, Hsp70, and Hsp40 system. MKT-077, JG-98, YM-1, and YM-8 are rhodocyanines that bind with low micromolar affinity to the nucleotide-binding domain of ADP- but not ATP-bound Hsp70, stabilizing the ADP-bound state (Pratt et al., 2015 ). The ADP-bound state of Hsp70 engages clients with higher affinity, and consequently MKT-077, JG-98, and YM-1 activate binding of Hsp70 to misfolded proteins (Wang et al., 2013 ; Pratt et al., 2015 ). Thus, under some conditions, these small molecules can promote folding of certain Hsp70 clients (Morishima et al., 2011 ; Pratt et al., 2015 ). However, prolonged interaction of clients with Hsp70 promotes their CHIP-dependent ubiquitylation and degradation in vivo (Morishima et al., 2011 ; Wang et al., 2013 ; Pratt et al., 2015 ). Intriguingly, YM-1 promotes clearance of polyglutamine oligomers and aggregates in cells (Wang et al., 2013 ; Pratt et al., 2015 ). MKT-0777, YM-1, JG-98, and YM-8 also promote clearance of tau and confer therapeutic benefit in tauopathy models (Abisambra et al., 2013 ; Miyata et al., 2013 ; Fontaine et al., 2015 ). Of importance, YM-8 is long lived in vivo and crosses the blood–brain barrier (Miyata et al., 2013 ). The dihydropyrimidine 115-7c activates Hsp70 ATPase turnover rate, promotes Hsp70 substrate refolding, and reduces α-synuclein aggregation in cell culture (Wisen et al., 2010 ; Kilpatrick et al., 2013 ). It binds to the IIA subdomain of Hsp70 and promotes the active Hsp70–Hsp40 complex (Wisen et al., 2010 ). Small-molecule enhancers of HtrA1 protease activity have also emerged (Jo et al., 2014 ). Thus it will be important to assess whether these small molecules enhance the activity of their respective disaggregases against various neurodegenerative substrates.Although small molecules that enhance disaggregase activity of endogenous human proteins might be the most immediately translatable, gene-, mRNA-, or protein-based therapies can also be envisioned. For example, adeno-associated viruses expressing enhanced disaggregases might be used to target degenerating neurons (Dong et al., 2005 ; Lo Bianco et al., 2008 ; Deverman et al., 2016 ). Alternatively, if viral vectors are undesirable, modified mRNAs might serve as an alternative to DNA-based gene therapy (Kormann et al., 2011 ). Protein-based therapeutics could also be explored. For example, intraperitoneal injection of human Hsp70 increased lifespan, delayed symptom onset, preserved motor function, and prolonged motor neuron viability in a mouse model of ALS (Gifondorwa et al., 2007 ; Gifondorwa et al., 2012 ). Several other studies suggest that exogenous delivery of Hsp70 can have beneficial, neuroprotective effects in mice (Nagel et al., 2008 ; Bobkova et al., 2014 ; Bobkova et al., 2015 ).Ultimately, if safety and ethical concerns can be overcome in a circumspect, risk-averse manner, CRISPR-Cas9–based therapeutics might even be used to genetically alter the underlying disaggregase to a potentiated form in selectively vulnerable neuronal populations. This approach might be particularly valuable if enhanced disaggregase activity is not detrimental in the long term. Moreover, stem cell–based therapies for replacing lost neurons could also be fortified to express enhanced disaggregase systems. Thus they would be endowed with resistance to potential infection by prion-like conformers that might have accumulated during disease progression (Cushman et al., 2010 ).Enhanced disaggregase activity is likely to be highly advantageous to neurons under circumstances in which protein misfolding has overwhelmed the system (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ). However, inappropriate hyperactivity of protein disaggregases might also have detrimental, off-target effects under regular conditions in which protein misfolding is not an overwhelming issue (Jackrel et al., 2014a ; Jackrel and Shorter, 2015 ). Thus it may be advantageous to engineer enhanced protein disaggregases to be highly substrate specific. In this way, off-target effects would be readily avoided. There is strong precedent for directed evolution or engineering of specialized chaperone or protein activity from a generalist antecedent (Wang et al., 2002 ; Farrell et al., 2007 ; Smith et al., 2015 ). Thus, engineering specialist disaggregases for each disease substrate could be achieved. Alternatively, transient or intermittent doses of enhanced disaggregases at specific times or places where they are most needed would also minimize potentially toxic side effects. For example, enhanced disaggregase activity might be applied ephemerally to clear existing misfolded conformers and then be withdrawn once the endogenous proteostasis network regains control. Similarly, it is straightforward to envision administration of small-molecule enhancers of disaggregase activity in intermittent protocols that enable facile recovery from potential side effects (Fontaine et al., 2015 ). In this way, any adverse effects of enhanced protein-disaggregase activity under normal physiological conditions would be avoided. Many barriers will need to be safely overcome to implement a successful therapeutic disaggregase, including how to deliver enhanced disaggregase activity to exactly where it is needed. However, these obstacles are not a reason to be pessimistic. On the contrary, the isolation of engineered disaggregases that efficaciously reverse deleterious misfolding of neurodegenerative disease proteins directs our attention to considerably expand the environs in which they should be sought. My closing sentences, therefore, are intended to be provocative.I suspect that neuroprotection could be broadly actualized via precise but subtle alterations to existing protein-disaggregase modalities. The engineering and evolution of protein disaggregases could yield important solutions to avert an imminent plague of neurodegenerative disorders that promises to devastate our society. I strongly suspect that cures for various neurodegenerative disorders will be realized by pioneering as-yet-uncharted regions of disaggregase sequence space or chemical space to elucidate small-molecule enhancers of disaggregase activity.  相似文献   

2.
l-Methionine γ-lyase (EC 4.4.1.11) was purified to homogeneity from Brevibacterium linens BL2, a coryneform bacterium which has been used successfully as an adjunct bacterium to improve the flavor of Cheddar cheese. The enzyme catalyzes the α,γ elimination of methionine to produce methanethiol, α-ketobutyrate, and ammonia. It is a pyridoxal phosphate-dependent enzyme, with a native molecular mass of approximately 170 kDa, consisting of four identical subunits of 43 kDa each. The purified enzyme had optimum activity at pH 7.5 and was stable at pHs ranging from 6.0 to 8.0 for 24 h. The pure enzyme had its highest activity at 25°C but was active between 5 and 50°C. Activity was inhibited by carbonyl reagents, completely inactivated by dl-propargylglycine, and unaffected by metal-chelating agents. The pure enzyme had catalytic properties similar to those of l-methionine γ-lyase from Pseudomonas putida. Its Km for the catalysis of methionine was 6.12 mM, and its maximum rate of catalysis was 7.0 μmol min−1 mg−1. The enzyme was active under salt and pH conditions found in ripening Cheddar cheese but susceptible to degradation by intracellular proteases.

Methanethiol is associated with desirable Cheddar-type sulfur notes in good-quality Cheddar cheese (2, 27). The mechanism for the production of methanethiol in cheese is unknown, but it is linked to the catabolism of methionine (1, 15). l-Methionine γ-lyase (EC 4.4.1.11; MGL), also known as methionase, l-methionine γ-demethiolase, and l-methionine methanethiollyase (deaminating), is a pyridoxal phosphate (PLP)-dependent enzyme that catalyzes the direct conversion of l-methionine to α-ketobutyrate, methanethiol, and ammonia by an α,γ-elimination reaction (26). It does not catalyze the conversion of d enantiomers (2426). MGL in Pseudomonas putida is a multifunctional enzyme system since it catalyzes the α,γ- and α,β-elimination reactions of methionine and its derivatives (24). In addition, the enzyme also catalyzes the β-replacement reactions of sulfur amino acids (24). Since its discovery in Escherichia coli and Proteus vulgaris by Onitake (19), this enzyme has been found in various bacteria and is regarded as a key enzyme in the bacterial metabolism of methionine. However, this enzyme has not been purified to homogeneity from any food-grade microorganisms.MGL is widely distributed in bacteria, especially in pseudomonads, and is induced by the addition of l-methionine to the culture medium (9, 28). The enzyme has been purified from Pseudomonas putida (25), Aeromonas sp. (26), Clostridium sporogenes (11), and Trichomonas vaginalis (16) and partially purified from and characterized for Brevibacterium linens NCDO 739 (4).B. linens is a nonmotile, non-spore-forming, non-acid-fast, gram-positive coryneform bacterium normally found on the surfaces of Limburger and other Trappist-type cheeses. This organism tolerates salt concentrations ranging between 8 and 20% and is capable of growing in a broad pH range from 5.5 to 9.5, with an optimum pH of 7.0 (20). In Trappist-type cheeses, brevibacteria depend on Saccharomyces cerevisiae to metabolize lactate, which increases the pH of the curd, as well as to produce growth factors that are important for their growth (20). Interest in B. linens has focused around its ability to produce an extracellular protease, which has recently been isolated (21), and its ability to produce high levels of methanethiol (3, 9, 10, 22).B. linens produces various sulfur compounds, including methanethiol, that are thought to be important in Cheddar-like flavor and aroma (3, 9, 10, 22). Ferchichi et al. (9) suggested that MGL is responsible for the methanethiol-producing capability of B. linens but did not provide definitive evidence. Weimer et al. (28) proposed that B. linens BL2 is responsible for Cheddar-type flavor development in low-fat cheese, but again conclusive evidence was lacking. In this study, MGL was purified to homogeneity from B. linens BL2 and its physical and chemical properties were examined.  相似文献   

3.
4.
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   

5.
6.
Insulin-dependent translocation of glucose transporter 4 (Glut4) to the plasma membrane of fat and skeletal muscle cells plays the key role in postprandial clearance of blood glucose. Glut4 represents the major cell-specific component of the insulin-responsive vesicles (IRVs). It is not clear, however, whether the presence of Glut4 in the IRVs is essential for their ability to respond to insulin stimulation. We prepared two lines of 3T3-L1 cells with low and high expression of myc7-Glut4 and studied its translocation to the plasma membrane upon insulin stimulation, using fluorescence-assisted cell sorting and cell surface biotinylation. In undifferentiated 3T3-L1 preadipocytes, translocation of myc7-Glut4 was low regardless of its expression levels. Coexpression of sortilin increased targeting of myc7-Glut4 to the IRVs, and its insulin responsiveness rose to the maximal levels observed in fully differentiated adipocytes. Sortilin ectopically expressed in undifferentiated cells was translocated to the plasma membrane regardless of the presence or absence of myc7-Glut4. AS160/TBC1D4 is expressed at low levels in preadipocytes but is induced in differentiation and provides an additional mechanism for the intracellular retention and insulin-stimulated release of Glut4.Adipocytes, skeletal muscle cells, and some neurons respond to insulin stimulation by translocating intracellular glucose transporter 4 (Glut4) to the plasma membrane. In all these cells, the insulin-responsive pool of Glut4 is localized in small membrane vesicles, the insulin-responsive vesicles (IRVs; Kandror and Pilch, 2011 ; Bogan, 2012 ). The protein composition of these vesicles has been largely characterized (Kandror and Pilch, 2011 ; Bogan, 2012 ). The IRVs consist predominantly of Glut4, insulin-responsive aminopeptidase (IRAP), sortilin, low-density-lipoprotein receptor–related protein 1 (LRP1), SCAMPs, and VAMP2. Glut4, IRAP, and sortilin physically interact with each other, which might be important for the biogenesis of the IRVs (Shi and Kandror, 2007 ; Shi et al., 2008 ). In addition, the IRVs compartmentalize recycling receptors, such as the transferrin receptor and the IGF2/mannose 6-phosphate receptor, although it is not clear whether these receptors represent obligatory vesicular components or their presence in the IRVs is explained by mass action (Pilch, 2008 ), inefficient sorting, or other reasons.Deciphering of the protein composition of the IRVs is important because it is likely to explain their unique functional property: translocation to the plasma membrane in response to insulin stimulation. Even if we presume that IRV trafficking is controlled by loosely associated peripheral membrane proteins, the latter should still somehow recognize the core vesicular components that create the “biochemical individuality” of this compartment. In spite of our knowledge of the IRV protein composition, however, the identity of the protein(s) that confer insulin sensitivity to these vesicles is unknown.Insulin responsiveness of the IRVs was associated with either IRAP or Glut4. Thus it was shown that Glut4 interacted with the intracellular anchor TUG (Bogan et al., 2003 , 2012 ), whereas IRAP associated with other proteins implemented in the regulation of Glut4 translocation, such as AS160 (Larance et al., 2005 ; Peck et al., 2006 ), p115 (Hosaka et al., 2005 ), tankyrase (Yeh et al., 2007 ), and several others (reviewed in Bogan, 2012 ). Results of these studies, or at least their interpretations, are not necessarily consistent with each other, as the existence of multiple independent anchors for the IRVs is, although possible, unlikely.Ablation of the individual IRV proteins has also led to controversial data. Thus knockout of IRAP decreases total protein levels of Glut4 but does not affect its translocation in the mouse model (Keller et al., 2002 ). On the contrary, knockdown of IRAP in 3T3-L1 adipocytes has a strong inhibitory effect on translocation of Glut4 (Yeh et al., 2007 ). In yet another study, knockdown of IRAP in 3T3-L1 adipocytes did not affect insulin-stimulated translocation of Glut4 but increased its plasma membrane content under basal conditions (Jordens et al., 2010 ). By the same token, total or partial ablation of Glut4 had various effects on expression levels, intracellular localization, and translocation of IRAP (Jiang et al., 2001 ; Abel et al., 2004 ; Carvalho et al., 2004 ; Gross et al., 2004 ; Yeh et al., 2007 ). Knockdown of either sortilin or LRP1 decreased protein levels of Glut4 (Shi and Kandror, 2005 ; Jedrychowski et al., 2010 ).One model that might explain these complicated and somewhat inconsistent results is that depletion of either major integral protein of the IRVs disrupts the network of interactions between vesicular proteins and thus decreases the efficiency of protein sorting into the IRVs (Kandror and Pilch, 2011 ). Correspondingly, the remaining IRV components that cannot be faithfully compartmentalized in the vesicles are either degraded (Jiang et al., 2001 ; Keller et al., 2002 ; Abel et al., 2004 ; Carvalho et al., 2004 ; Shi and Kandror, 2005 ; Yeh et al., 2007 ; Jedrychowski et al., 2010 ) or mistargeted (Jiang et al., 2001 ; Jordens et al., 2010 ), depending on experimental conditions and types of cells used in these studies. In other words, knockdown of any major IRV component may decrease vesicle formation along with insulin responsiveness. Thus, in spite of a large body of literature, the identity of protein(s) that confer insulin responsiveness to the IRVs is unknown.Here we used a gain-of-function approach to address this question. Specifically, we attempted to “build” functional IRVs in undifferentiated 3T3-L1 preadipocytes by forced expression of the relevant proteins. Undifferentiated preadipocytes do not express Glut4 or sortilin and lack IRVs (ElJack et al., 1999 ; Shi and Kandror, 2005 ; Shi et al., 2008 ). Correspondingly, IRAP, which is expressed in these cells, shows low insulin response (Ross et al., 1998 ; Shi et al., 2008 ). We found that ectopic expression of increasing amounts of Glut4 in undifferentiated preadipocytes does not lead to its marked translocation to the plasma membrane upon insulin stimulation. On the contrary, sortilin expressed in undifferentiated preadipocytes was localized in the IRVs and was translocated to the plasma membrane in response to insulin stimulation. Moreover, upon coexpression with Glut4, sortilin dramatically increased its insulin responsiveness to the levels observed in fully differentiated adipocytes. Thus sortilin may represent the key component of the IRVs, which is responsible not only for the formation of vesicles (Shi and Kandror, 2005 ; Ariga et al., 2008 ; Hatakeyama and Kanzaki, 2011 ), but also for their insulin responsiveness. It is worth noting that sortilin levels are significantly decreased in obese and diabetic humans and mice (Kaddai et al., 2009 ). We thus suggest that sortilin may be a novel and important target in the fight against insulin resistance and diabetes.Our experiments also demonstrate that undifferentiated preadipocytes lack a mechanism for the full intracellular retention of Glut4 that can be achieved by ectopic expression of AS160/TBC1D4.  相似文献   

7.
8.
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   

9.
A low-specificity l-threonine aldolase (l-TA) gene from Pseudomonas sp. strain NCIMB 10558 was cloned and sequenced. The gene contains an open reading frame consisting of 1,041 nucleotides corresponding to 346 amino acid residues. The gene was overexpressed in Escherichia coli cells, and the recombinant enzyme was purified and characterized. The enzyme, requiring pyridoxal 5′-phosphate as a coenzyme, is strictly l specific at the α position, whereas it cannot distinguish between threo and erythro forms at the β position. In addition to threonine, the enzyme also acts on various other l-β-hydroxy-α-amino acids, including l-β-3,4-dihydroxyphenylserine, l-β-3,4-methylenedioxyphenylserine, and l-β-phenylserine. The predicted amino acid sequence displayed less than 20% identity with those of low-specificity l-TA from Saccharomyces cerevisiae, l-allo-threonine aldolase from Aeromonas jandaei, and four relevant hypothetical proteins from other microorganisms. However, lysine 207 of low-specificity l-TA from Pseudomonas sp. strain NCIMB 10558 was found to be completely conserved in these proteins. Site-directed mutagenesis experiments showed that substitution of Lys207 with Ala or Arg resulted in a significant loss of enzyme activity, with the corresponding disappearance of the absorption maximum at 420 nm. Thus, Lys207 of the l-TA probably functions as an essential catalytic residue, forming an internal Schiff base with the pyridoxal 5′-phosphate of the enzyme to catalyze the reversible aldol reaction.β-Hydroxy-α-amino acids constitute an important class of compounds. They are natural products in their own right and are components of a range of antibiotics, for example, cyclosporin A, lysobactin, and vancomycin (30) and bouvardin and deoxybouvardin (6). 4-Hydroxy-l-threonine is a precursor of rizobitoxine, a potent inhibitor of pyridoxal 5′-phosphate (PLP)-dependent enzymes (32). 3,4,5-Trihydroxyl-l-aminopentanoic acid is a key component of polyoxins (32). l-threo-3,4-Dihydroxyphenylserine is a new drug for Parkinson’s disease therapy (13). However, the industrial production of β-hydroxy-α-amino acids has been limited to chemical synthesis processes, which need multiple steps to isolate the four isomers (l-threo form, d-threo form, l-erythro form, and d-erythro form). Threonine aldolase (EC 4.1.2.5), which stereospecifically catalyzes the retro-aldol cleavage of threonine, is a potentially useful catalyst for the synthesis of substituted amino acids from aldehyde and glycine (27, 31, 32).Two different types of threonine aldolases are known so far. l-allo-Threonine aldolase (l-allo-TA), isolated and purified from Aeromonas jandaei DK-39 (8), stereospecifically catalyzes the reversible interconversion of l-allo-threonine and glycine. Low-specificity l-threonine aldolase (l-TA) catalyzes the cleavage of both l-threonine and l-allo-threonine to glycine and acetaldehyde, as well as the reverse reaction, aldol condensation. The enzymes have been purified and characterized from Candida humicola (9, 34) and Saccharomyces cerevisiae (12). Low-specificity l-TA activity has also been shown to exist in mammals (7, 23, 26) and a variety of other microbial species (2, 4, 17, 35). The enzyme is physiologically important for the synthesis of cellular glycine in yeast (12, 15, 16). Threonine aldolases with distinct stereospecificities are ideal targets for enzymology studies on structural and functional relationships. However, information on the primary structures of threonine aldolases was limited to our recent studies (11, 12). The construction of an overproduction system for threonine aldolase will be indispensable for the industrial biosyntheses of β-hydroxy-α-amino acids.The present work focuses on the cloning, sequencing, and overexpression in Escherichia coli cells of the low-specificity l-TA gene from Pseudomonas sp. strain NCIMB 10558, the purification and characterization of the recombinant enzyme, and the identification of the active-site lysine residue of the enzyme by site-directed mutagenesis. Evidence is presented that Lys207 of low-specificity l-TA probably functions as a catalytic residue, forming an internal Schiff base with the PLP of the enzyme to catalyze the reversible aldol reaction. This is the first report showing a purified enzyme with l-β-3,4-dihydroxyphenylserine aldolase and l-β-3,4-methylenedioxyphenylserine aldolase activities, providing a new route for the industrial production of these important unnatural amino acids.  相似文献   

10.
Two methods were used to compare the biodegradation of six polychlorinated biphenyl (PCB) congeners by 12 white rot fungi. Four fungi were found to be more active than Phanerochaete chrysosporium ATCC 24725. Biodegradation of the following congeners was monitored by gas chromatography: 2,3-dichlorobiphenyl, 4,4′-dichlorobiphenyl, 2,4′,5-trichlorobiphenyl (2,4′,5-TCB), 2,2′,4,4′-tetrachlorobiphenyl, 2,2′,5,5′-tetrachlorobiphenyl, and 2,2′,4,4′,5,5′-hexachlorobiphenyl. The congener tested for mineralization was 2,4′,5-[U-14C]TCB. Culture supernatants were also assayed for lignin peroxidase and manganese peroxidase activities. Of the fungi tested, two strains of Bjerkandera adusta (UAMH 8258 and UAMH 7308), one strain of Pleurotus ostreatus (UAMH 7964), and Trametes versicolor UAMH 8272 gave the highest biodegradation and mineralization. P. chrysosporium ATCC 24725, a strain frequently used in studies of PCB degradation, gave the lowest mineralization and biodegradation activities of the 12 fungi reported here. Low but detectable levels of lignin peroxidase and manganese peroxidase activity were present in culture supernatants, but no correlation was observed among any combination of PCB congener biodegradation, mineralization, and lignin peroxidase or manganese peroxidase activity. With the exception of P. chrysosporium, congener loss ranged from 40 to 96%; however, these values varied due to nonspecific congener binding to fungal biomass and glassware. Mineralization was much lower, ≤11%, because it measures a complete oxidation of at least part of the congener molecule but the results were more consistent and therefore more reliable in assessment of PCB biodegradation.

Polychlorinated biphenyls (PCBs) are produced by chlorination of biphenyl, resulting in up to 209 different congeners. Commercial mixtures range from light oily fluids to waxes, and their physical properties make them useful as heat transfer fluids, hydraulic fluids, solvent extenders, plasticizers, flame retardants, organic diluents, and dielectric fluids (1, 21). Approximately 24 million lb are in the North American environment (19). The stability and hydrophobic nature of these compounds make them a persistent environmental hazard.To date, bacterial transformations have been the main focus of PCB degradation research. Aerobic bacteria use a biphenyl-induced dioxygenase enzyme system to attack less-chlorinated congeners (mono- to hexachlorobiphenyls) (1, 5, 7, 8, 22). Although more-chlorinated congeners are recalcitrant to aerobic bacterial degradation, microorganisms in anaerobic river sediments reductively dechlorinate these compounds, mainly removing the meta and para chlorines (1, 6, 10, 33, 34).The degradation of PCBs by white rot fungi has been known since 1985 (11, 18). Many fungi have been tested for their ability to degrade PCBs, including the white rot fungi Coriolus versicolor (18), Coriolopsis polysona (41), Funalia gallica (18), Hirneola nigricans (35), Lentinus edodes (35), Phanerochaete chrysosporium (3, 11, 14, 17, 18, 35, 39, 4143), Phlebia brevispora (18), Pleurotus ostreatus (35, 43), Poria cinerescens (18), Px strain (possibly Lentinus tigrinus) (35), and Trametes versicolor (41, 43). There have also been studies of PCB metabolism by ectomycorrhizal fungi (17) and other fungi such as Aspergillus flavus (32), Aspergillus niger (15), Aureobasidium pullulans (18), Candida boidinii (35), Candida lipolytica (35), Cunninghamella elegans (16), and Saccharomyces cerevisiae (18, 38). The mechanism of PCB biodegradation has not been definitively determined for any fungi. White rot fungi produce several nonspecific extracellular enzymes which have been the subject of extensive research. These nonspecific peroxidases are normally involved in lignin degradation but can oxidize a wide range of aromatic compounds including polycyclic aromatic hydrocarbons (37). Two peroxidases, lignin peroxidase (LiP) and Mn peroxidase (MnP), are secreted into the environment of the fungus under conditions of nitrogen limitation in P. chrysosporium (23, 25, 27, 29) but are not stress related in fungi such as Bjerkandera adusta or T. versicolor (12, 30).Two approaches have been used to determine the biodegradability of PCBs by fungi: (i) loss of the parent congener analyzed by gas chromatography (GC) (17, 32, 35, 42, 43) and (ii) mineralization experiments in which the 14C of the universally labeled 14C parent congener is recovered as 14CO2 (11, 14, 18, 39, 41). In the first method, the loss of a peak on a chromatogram makes it difficult to decide whether the PCB is being partly degraded, mineralized, adsorbed to the fungal biomass, or bound to glassware, soil particles, or wood chips. Even when experiments with killed-cell and abiotic controls are performed, the extraction efficiency and standard error can make data difficult to interpret. For example, recoveries can range anywhere from 40 to 100% depending on the congener used and the fungus being investigated (17). On the other hand, recovery of significant amounts of 14CO2 from the cultures incubated with a 14C substrate provides definitive proof of fungal metabolism. There appears to be only one report relating data from these two techniques (18), and in that study, [U-14C]Aroclor 1254, rather than an individual congener, was used.In this study, we examined the ability of 12 white rot fungal strains to metabolize selected PCB congeners to determine which strains were the most active degraders. Included in this group was P. chrysosporium ATCC 24725, a strain used extensively in PCB studies (3, 14, 18, 35, 39, 4143). Six PCB congeners were selected to give a range of chlorine substitutions and therefore a range of potential biodegradability which was monitored by GC. One of the chosen congeners was 14C labeled and used in studies to compare the results from a mineralization method with those from the GC method.  相似文献   

11.
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   

12.
SPA2 encodes a yeast protein that is one of the first proteins to localize to sites of polarized growth, such as the shmoo tip and the incipient bud. The dynamics and requirements for Spa2p localization in living cells are examined using Spa2p green fluorescent protein fusions. Spa2p localizes to one edge of unbudded cells and subsequently is observable in the bud tip. Finally, during cytokinesis Spa2p is present as a ring at the mother–daughter bud neck. The bud emergence mutants bem1 and bem2 and mutants defective in the septins do not affect Spa2p localization to the bud tip. Strikingly, a small domain of Spa2p comprised of 150 amino acids is necessary and sufficient for localization to sites of polarized growth. This localization domain and the amino terminus of Spa2p are essential for its function in mating. Searching the yeast genome database revealed a previously uncharacterized protein which we name, Sph1p (Spa2p homolog), with significant homology to the localization domain and amino terminus of Spa2p. This protein also localizes to sites of polarized growth in budding and mating cells. SPH1, which is similar to SPA2, is required for bipolar budding and plays a role in shmoo formation. Overexpression of either Spa2p or Sph1p can block the localization of either protein fused to green fluorescent protein, suggesting that both Spa2p and Sph1p bind to and are localized by the same component. The identification of a 150–amino acid domain necessary and sufficient for localization of Spa2p to sites of polarized growth and the existence of this domain in another yeast protein Sph1p suggest that the early localization of these proteins may be mediated by a receptor that recognizes this small domain.Polarized cell growth and division are essential cellular processes that play a crucial role in the development of eukaryotic organisms. Cell fate can be determined by cell asymmetry during cell division (Horvitz and Herskowitz, 1992; Cohen and Hyman, 1994; Rhyu and Knoblich, 1995). Consequently, the molecules involved in the generation and maintenance of cell asymmetry are important in the process of cell fate determination. Polarized growth can occur in response to external signals such as growth towards a nutrient (Rodriguez-Boulan and Nelson, 1989; Eaton and Simons, 1995) or hormone (Jackson and Hartwell, 1990a , b ; Segall, 1993; Keynes and Cook, 1995) and in response to internal signals as in Caenorhabditis elegans (Goldstein et al., 1993; Kimble, 1994; Priess, 1994) and Drosophila melanogaster (St Johnston and Nusslein-Volhard, 1992; Anderson, 1995) early development. Saccharomyces cerevisiae undergo polarized growth towards an external cue during mating and to an internal cue during budding. Polarization towards a mating partner (shmoo formation) and towards a new bud site requires a number of proteins (Chenevert, 1994; Chant, 1996; Drubin and Nelson, 1996). Many of these proteins are necessary for both processes and are localized to sites of polarized growth, identified by the insertion of new cell wall material (Tkacz and Lampen, 1972; Farkas et al., 1974; Lew and Reed, 1993) to the shmoo tip, bud tip, and mother–daughter bud neck. In yeast, proteins localized to growth sites include cytoskeletal proteins (Adams and Pringle, 1984; Kilmartin and Adams, 1984; Ford, S.K., and J.R. Pringle. 1986. Yeast. 2:S114; Drubin et al., 1988; Snyder, 1989; Snyder et al., 1991; Amatruda and Cooper, 1992; Lew and Reed, 1993; Waddle et al., 1996), neck filament components (septins) (Byers and Goetsch, 1976; Kim et al., 1991; Ford and Pringle, 1991; Haarer and Pringle, 1987; Longtine et al., 1996), motor proteins (Lillie and Brown, 1994), G-proteins (Ziman, 1993; Yamochi et al., 1994; Qadota et al., 1996), and two membrane proteins (Halme et al., 1996; Roemer et al., 1996; Qadota et al., 1996). Septins, actin, and actin-associated proteins localize early in the cell cycle, before a bud or shmoo tip is recognizable. How this group of proteins is localized to and maintained at sites of cell growth remains unclear.Spa2p is one of the first proteins involved in bud formation to localize to the incipient bud site before a bud is recognizable (Snyder, 1989; Snyder et al., 1991; Chant, 1996). Spa2p has been localized to where a new bud will form at approximately the same time as actin patches concentrate at this region (Snyder et al., 1991). An understanding of how Spa2p localizes to incipient bud sites will shed light on the very early stages of cell polarization. Later in the cell cycle, Spa2p is also found at the mother–daughter bud neck in cells undergoing cytokinesis. Spa2p, a nonessential protein, has been shown to be involved in bud site selection (Snyder, 1989; Zahner et al., 1996), shmoo formation (Gehrung and Snyder, 1990), and mating (Gehrung and Snyder, 1990; Chenevert et al., 1994; Yorihuzi and Ohsumi, 1994; Dorer et al., 1995). Genetic studies also suggest that Spa2p has a role in cytokinesis (Flescher et al., 1993), yet little is known about how this protein is localized to sites of polarized growth.We have used Spa2p green fluorescent protein (GFP)1 fusions to investigate the early localization of Spa2p to sites of polarized growth in living cells. Our results demonstrate that a small domain of ∼150 amino acids of this large 1,466-residue protein is sufficient for targeting to sites of polarized growth and is necessary for Spa2p function. Furthermore, we have identified and characterized a novel yeast protein, Sph1p, which has homology to both the Spa2p amino terminus and the Spa2p localization domain. Sph1p localizes to similar regions of polarized growth and sph1 mutants have similar phenotypes as spa2 mutants.  相似文献   

13.
Acyl coenzyme A (CoA) synthetase (EC 6.2.1.8) from Pseudomonas fragi catalyzes the synthesis of adenosine 5′-tetraphosphate (p4A) and adenosine 5′-pentaphosphate (p5A) from ATP and tri- or tetrapolyphosphate, respectively. dATP, adenosine-5′-O-[γ-thiotriphosphate] (ATPγS), adenosine(5′)tetraphospho(5′)adenosine (Ap4A), and adenosine(5′)pentaphospho(5′)adenosine (Ap5A) are also substrates of the reaction yielding p4(d)A in the presence of tripolyphosphate (P3). UTP, CTP, and AMP are not substrates of the reaction. The Km values for ATP and P3 are 0.015 and 1.3 mM, respectively. Maximum velocity was obtained in the presence of MgCl2 or CoCl2 equimolecular with the sum of ATP and P3. The relative rates of synthesis of p4A with divalent cations were Mg = Co > Mn = Zn >> Ca. In the pH range used, maximum and minimum activities were measured at pH values of 5.5 and 8.2, respectively; the opposite was observed for the synthesis of palmitoyl-CoA, with maximum activity in the alkaline range. The relative rates of synthesis of palmitoyl-CoA and p4A are around 10 (at pH 5.5) and around 200 (at pH 8.2). The synthesis of p4A is inhibited by CoA, and the inhibitory effect of CoA can be counteracted by fatty acids. To a lesser extent, the enzyme catalyzes the synthesis also of Ap4A (from ATP), Ap5A (from p4A), and adenosine(5′)tetraphospho(5′)nucleoside (Ap4N) from adequate adenylyl donors (ATP, ATPγS, or octanoyl-AMP) and adequate adenylyl acceptors (nucleoside triphosphates).Dinucleoside polyphosphates have been detected in a wide variety of eukaryotic and prokaryotic organisms (13). In higher organisms, their concentrations are generally on the order of 0.01 to 1 μM. Human blood platelets and chromaffin cells of bovine adrenal medulla contain diadenosine polyphosphates located in the dense bodies (10, 26, 35) and chromaffin granules (32, 38), respectively, where they may reach higher local concentrations. The occurrence of dinucleoside polyphosphates has been described for lower eukaryotic (Saccharomyces cerevisiae, Dictyostelium discoideum, and Physarum polycephalum) and for prokaryotic (Salmonella typhimurium, Escherichia coli, and Clostridium acetobutylicum) organisms (13).Dinucleoside tetraphosphates participate in the control of purine nucleotide metabolism (36), where Ap4A is an activator of both the IMP-GMP-specific cytosolic 5′-nucleotidase (EC 3.1.3.5) and AMP deaminase (EC 3.5.4.6) (Ka, micromolar range) and Gp4G is an activator of GMP reductase (EC 1.6.6.8) (Ka, nanomolar range) (36). As the concentration of dinucleoside polyphosphates increases under unfavorable environmental conditions, they have been implicated in the cellular response to stress (31). A role of Ap4A in DNA synthesis has been proposed elsewhere (14). Dinucleoside polyphosphates are also transition state analogs of some kinases (37). More recently, the dinucleoside triphosphatase activity of a putative tumor suppressor gene product has been described (3).The nucleoside 5′-polyphosphates (pnN) are another family of related compounds, p4A has been detected in rabbit and horse muscle (41), rat liver (44), S. cerevisiae spores (19), and chromaffin granules (38). As p4A is a very strong inhibitor (Ki, nanomolar range) of asymmetrical dinucleoside tetraphosphatase (EC 3.6.1.17) (22), changes in the level of p4A could affect the concentration and physiological roles of Ap4A. Other enzymes known to be inhibited (Ki, micromolar range) by p4N are guanylate cyclase (EC 4.6.1.2) (p4A and p4G) (18) and phosphodiesterase I (EC 3.1.4.1) (p4G) (9). Effects of p4A on the tone of the vascular system, mediated by P2 receptors, have also been described elsewhere (21).The cellular level of dinucleoside polyphosphates results from their rate of degradation and synthesis. The following specific enzymes, implicated in the cleavage of dinucleoside polyphosphates, have been described (see reference 15 for a review): asymmetrical dinucleoside tetraphosphatase (EC 3.6.1.17), symmetrical dinucleoside tetraphosphatase (EC 3.6.1.41), dinucleoside tetraphosphate phosphorylase (EC 2.7.7.53), and dinucleoside triphosphatase (EC 3.6.1.29). In addition, there are other unspecific enzymes able to catalyze the hydrolysis of dinucleoside polyphosphates like E. coli 5′-nucleotidase (34) and phosphodiesterase I (9, 15, 26).This paper deals with the synthesis of (di)nucleoside polyphosphates. It has been known since 1966 that some aminoacyl tRNA synthetases (30, 45) catalyze the synthesis of Ap4A through reactions 1 and 2: reaction 1 reaction 2 The possibility that other enzymes (mainly synthetases and some transferases) which catalyze the formation of AMP, via nucleotidyl-containing intermediates and by releasing PPi, could catalyze the synthesis of dinucleoside polyphosphates was later raised (17). Luciferase (EC 1.13.12.7), considered as an oxidoreductase, catalyzes the synthesis of Ap4A with ATP as substrate and luciferin as an essential activator (27, 40): reaction 3 reaction 4 Acetyl-CoA synthetase (EC 6.2.1.1) from S. cerevisiae also catalyzes the synthesis of p4A and p5A, from ATP and P3 and P4, respectively (16). In the reactions catalyzed by luciferase and acetyl-CoA synthetase, ATP is a very good substrate for the formation of the E · X-AMP complex (X = the appropriate acyl residue), whereas any NTP (or even P3) is an acceptor (particularly in the case of luciferase) of the AMP moiety of the complex, provided that it has an intact terminal pyrophosphate (27, 40).Here we show that acyl-CoA synthetase from Pseudomonas fragi catalyzes the synthesis of p4A, p5A, Ap4A, Ap5A, and a variety of Ap4Ns. In our view, these findings widen the knowledge of the mechanisms of synthesis of (di)nucleoside polyphosphates in prokaryotes and, by extrapolation, also in eukaryotes.  相似文献   

14.
A 4,103-bp long DNA fragment containing the structural gene of a gentisate 1,2-dioxygenase (EC 1.13.11.4), gtdA, from Sphingomonas sp. strain RW5 was cloned and sequenced. The gtdA gene encodes a 350-amino-acid polypeptide with a predicted size of 38.85 kDa. Comparison of the gtdA gene product with protein sequences in databases, including those of intradiol or extradiol ring-cleaving dioxygenases, revealed no significant homology except for a low similarity (27%) to the 1-hydroxy-2-naphthoate dioxygenase (phdI) of the phenanthrene degradation in Nocardioides sp. strain KP7 (T. Iwabuchi and S. Harayama, J. Bacteriol. 179:6488–6494, 1997). This gentisate 1,2-dioxygenase is thus a member of a new class of ring-cleaving dioxygenases. The gene was subcloned and hyperexpressed in E. coli. The resulting product was purified to homogeneity and partially characterized. Under denaturing conditions, the polypeptide exhibited an approximate size of 38.5 kDa and migrated on gel filtration as a species with a molecular mass of 177 kDa. The enzyme thus appears to be a homotetrameric protein. The purified enzyme stoichiometrically converted gentisate to maleylpyruvate, which was identified by gas chromatography-mass spectrometry analysis as its methyl ester. Values of affinity constants (Km) and specificity constants (Kcat/Km) of the enzyme were determined to be 15 μM and 511 s−1 M−1 × 104 for gentisate and 754 μM and 20 s−1 M−1 × 104 for 3,6-dichlorogentisate. Three further open reading frames (ORFs) were found downstream of gtdA. The deduced amino acid sequence of ORF 2 showed homology to several isomerases and carboxylases, and those of ORFs 3 and 4 exhibited significant homology to enzymes of the glutathione isomerase superfamily and glutathione reductase superfamily, respectively.Large amounts of aromatic compounds have been released into the environment during the last decades as a result of extensive production of industrial chemicals and agricultural applications of pesticides. Many of these compounds, particularly the chlorinated derivatives, are toxic, even at low concentrations, and persist in the environment (14, 39). Numerous microorganisms have been isolated which degrade xenobiotic aromatic compounds through aerobic or anaerobic degradative reactions (16, 17, 34, 46, 47). A wide variety of polycyclic and homocyclic aromatic compounds are aerobically transformed to a limited number of central dihydroxylated intermediates like catechol, protocatechuate, or gentisate. Whereas catabolic pathways for catechol and protocatechuate have been extensively characterized (22, 35), little is known about gentisate degradation.Gentisic acid (2,5-dihydroxybenzoic acid) is a key intermediate in the aerobic degradation of such aromatic compounds as dibenzofuran (15), naphthalene (18, 48), salicylate (40, 55), anthranilate (32), and 3-hydroxybenzoate (26). Degradation of gentisate is initiated by gentisate 1,2-dioxygenase (GDO; EC 1.13.11.4, gentisate:oxygen oxidoreductase), which cleaves the aromatic ring between the carboxyl and the vicinal hydroxyl group to form maleylpyruvate (30). Maleylpyruvate can be converted to central metabolites of the Krebs cycle either by cleavage to pyruvate and maleate (5, 24) or by isomerization to fumarylpyruvate and subsequent cleavage to fumarate and pyruvate (10, 31, 51).Of the two well-studied classes of ring cleavage dioxygenases, intradiol enzymes, such as catechol 1,2-dioxygenase or protocatechuate 3,4-dioxygenase, contain an Fe3+ atom in the catalytic center and cleave the aromatic substrate between two vicinal hydroxyl groups (7, 37, 38), whereas dioxygenases of the extradiol class, such as catechol 2,3-dioxygenase or protocatechuate 4,5-dioxygenase, contain Fe2+ and cleave the aromatic substrate adjacent to two vicinal hydroxyl groups (1, 13). Gentisate 1,2-dioxygenase cleaves aromatic rings containing hydroxyl groups situated para to one another. Although the mechanism of oxygen activation was proposed to be similar to that of enzymes of the extradiol dioxygenase class (20), and the active center contains Fe2+ (11, 21, 29, 49, 51), the Fe2+ is not bound to the enzyme by electron-donating ligands such as cysteine or tyrosine (21) as is the case for extradiol-cleaving dioxygenases (19). It is being assumed, therefore, that GDO represents a novel class of ring-cleaving dioxygenases.GDOs have been purified and characterized from gram-positive bacteria of the genera Bacillus and Rhodococcus (29, 50) and gram-negative bacteria of the genera Klebsiella, Comamonas, and Moraxella (11, 21, 49), and amino-terminal sequences of GDOs from Comamonas testosteroni and Comamonas acidovorans have been determined (21), but until now, no complete sequence of any GDO or of a gene encoding GDO has been reported. Here we describe the cloning and sequencing of the gene encoding the GDO from Sphingomonas sp. strain RW5 and its partial characterization. This GDO represents a novel class of dioxygenases with very low similarity to any other known ring-cleaving dioxygenases.  相似文献   

15.
Whipple’s disease is a systemic disorder in which a gram-positive rod-shaped bacterium is constantly present in infected tissues. After numerous unsuccessful attempts to culture this bacterium, it was eventually characterized by 16S rRNA gene analysis to be a member of the actinomycetes. The name Tropheryma whippelii was proposed. Until now, the bacterium has only been found in infected human tissues, but there is no evidence for human-to-human transmission. Here we report the detection of DNA specific for the Whipple’s disease bacterium in 25 of 38 wastewater samples from five different sewage treatment plants in the area of Heidelberg, Germany. These findings provide the first evidence that T. whippelii occurs in the environment, within a polymicrobial community. This is in accordance with the phylogenetic relationship of this bacterium as well as with known epidemiological aspects of Whipple’s disease. Our data argue for an environmental source for infection with the Whipple’s disease bacterium.Whipple’s disease, which was first described in 1907 as intestinal lipodystrophy (18), is a multisystem disorder characterized by the presence of gram-positive, rod-shaped bacteria in infected human tissues (2). Numerous attempts to culture the bacterium associated with Whipple’s disease have failed (2), and eventually its phylogenetic position within the actinomycetes was established by 16S rRNA gene analysis (11, 19). The name Tropheryma whippelii was proposed (11).The clinical presentation of Whipple’s disease is heterogeneous. Frequently, patients suffer for years from arthralgias, chronic diarrhea, and weight loss, and less often from dementia or cardiac insufficiency. If untreated, the disease is usually fatal, but with appropriate antibiotic therapy the prognosis is favorable (2). However, the pathogenesis of Whipple’s disease appears to involve more than just an infection with T. whippelii. Immunological abnormalities are presumed to play a necessary role as predisposing factors (2, 7, 8), a view which is strengthened by the detection of T. whippelii in association with AIDS (5).Two outstanding questions in the epidemiology of Whipple’s disease are the bacterium’s natural habitat and the route of infection. Until now, T. whippelii has never been found outside of infected human hosts, and although an oral route of infection has often been suspected (10), it has not been proven. There is no evidence for human-to-human transmission.A reassessment of its phylogeny revealed a close relationship of the Whipple’s disease bacterium to typical environmental bacteria, such as the cellulomonads and the rare group B peptidoglycan organisms (6). This would support the hypothesis that T. whippelii may be a soil or water inhabitant. This may also explain the difficulties involved with its culture, as an estimated 80 to 99% of bacteria occurring in such natural environments are not culturable on artificial media (14, 17) and uncultured members of the actinomycete line of descent have been found in different environments and geographical locations (12). Soil and water bacteria tend to concentrate in sewage to form communities in which a large variety of different species are found. Therefore, the search for the Whipple’s disease bacterium was concentrated at sewage treatment plants.From September 1995 to July 1996, a total of 38 effluent samples from the sedimentation ponds of five different sewage treatment plants were examined by PCR for the presence of the Whipple’s disease bacterium. The sewage treatment plants are located within a distance of 5 to 25 km from Heidelberg, Germany (Fig. (Fig.1).1). Effluent was sampled in 1,000-ml single-use plastic flasks and filtered as described previously (4) in portions of 25 ml through cellulose acetate prefilters (25-mm diameter, 5-μm pore size; Sartorius, Göttingen, Germany). It was then passed through polyvinylidene fluoride filters (25-mm diameter, 0.45-μm pore size; Millipore, Bedford, Mass.). The DNA from the 0.45-μm-pore-size filters was extracted with the EnviroAmp Legionella sample preparation kit (Perkin-Elmer, Norwalk, Conn.), with the addition of 0.2% bovine serum albumin to the water in the final extraction step. Distilled water for negative controls was sampled in the same flasks and treated in the same way as the sewage. Open in a separate windowFIG. 1Map of the area around Heidelberg, Germany, displaying the rivers Rhine (Rhein) and Neckar, the relevant townships (shaded areas with names in capital letters), and the five sewage treatment plants (dots) from which samples were taken. The inset in the figure shows the localization of the area within Germany.PCR amplification and detection were performed as described in detail elsewhere (15). Primers whip1 (5′-AGAGATACGCCCCCCGCAA) and whip2 (5′-ATTCGCTCCACCTTGCGA), which amplify a fragment of 267 bp from the 16S rRNA gene of the bacterium, were used. In addition to hybridization with oligonucleotide whip3 (5′-TGGTACAGAGGGTTGCAATA), a second hybridization in a separate reaction was performed with oligonucleotide whip4 (5′-GTAATGGCGGGGACTCACAG). The specificities of primers whip1 and whip2 and of probe whip3 have been tested previously (15), and testing of probe whip4 with the same set of 37 control bacteria gave the same results as those reported for whip3 (15).Of the 38 sewage samples tested, 25 were found to be positive by PCR and hybridization with both oligonucleotides (Table (Table1).1). Positive samples were found at each of the five sewage treatment plants. Most of the positive PCR products displayed relatively weak bands at the expected size of 267 bp. They also contained some extraneous bands, which were not of the expected size, probably due to the large amount of DNA from other organisms present in sewage. On hybridization, however, the PCR products displayed distinct bands, which were consistent between the two oligonucleotides (Fig. (Fig.2).2).

TABLE 1

Results of PCR for the Whipple’s disease bacterium from sewage samples
Sewage treatment plantNo. of samples:
TestedPositive
Heidelberg149
Untere Hardt77
Mannheim137
Kohlhof21
Grenzhof21
 Total3825
Open in a separate windowOpen in a separate windowFIG. 2Detection of DNA from the Whipple’s disease bacterium in sewage samples. Shown are the results of analyses of PCR products by polyacrylamide gel electrophoresis (A) and Southern blot hybridization with oligonucleotide whip4 (B). Lanes: 1 and 10, 100-bp DNA marker; 2, positive control consisting of a tissue digest (diluted 10−4) from the intestinal biopsy of a patient with Whipple’s disease (9); 3, negative control consisting of 10 μl of distilled water in the PCR mixture; 4 and 5, sewage samples from Heidelberg; 6 and 9, negative controls consisting of distilled water which was filtered and treated in the same way as the sewage; 7 and 8, sewage samples from Untere Hardt.To further compare the sequences of the PCR products with the known sequence of the Whipple’s disease bacterium, nine PCR products selected from all five sewage treatment plants were cloned and sequenced. The PCR products were reamplified over 5 cycles by using specific primers with BamHI and EcoRI restriction endonuclease recognition sites (11), ligated into vector pDS1-NOC (3) and transformed by electroporation into Escherichia coli JM101. To screen for positive clones, a series of PCRs (primer whip2 and vector-specific primer 5′-TTGCTTTGTGAGCGGATAACAATTAT) in which pools of 10 E. coli colonies were amplified in each of the reactions was performed. The PCR products were tested by gel electrophoresis, Southern blotting, and hybridization with oligonucleotide whip4. Pools which produced hybridizing bands of the appropriate size were retested to find individual colonies. Plasmids from positive clones were extracted according to standard methods (13), and both strands of the inserts were sequenced manually with the AmpliCycle sequencing kit (Perkin-Elmer), with incorporation of [α-33P]dATP. For each of the nine sewage samples, the sequence of the PCR product was identical to the known sequence determined previously for the Whipple’s disease bacterium (5, 9).This is the first documented encounter with the Whipple’s disease bacterium outside of the human body. The specific sequence was found in all of the sewage treatment plants from which samples were taken, indicating that the bacterium is a regular member of, and most likely is able to multiply in, such polymicrobial communities. This finding is in agreement with one previous report which found sequences highly similar to, but not identical with, those of the Whipple’s disease bacterium in the eutrophic water of a Mediterranean coastal lagoon (1).The environmental occurrence of the Whipple’s disease bacterium at the sewage treatment plants is in agreement with the evolutionary relationships of the bacterium (6, 11). It is also consistent with the lack of geographical preference of reported cases of Whipple’s disease in Germany and with the relatively constant incidence of new cases per year (16). In addition, it is consistent with a predominance of outdoor professions among patients with the disease (2). These features and the absence of the Whipple’s disease bacterium in normal human tissues (11, 15) make it highly unlikely that this bacterium is a common commensal which is transferred between warm-blooded hosts in the same way as a variety of other pathogens. However, the complete range of occurrences in the environment and the exact localization of the habitats of the Whipple’s disease bacterium still remain to be determined.

Nucleotide sequence accession numbers.

The GenBank/EMBL accession numbers of the published 16S rRNA gene sequences of the Whipple’s disease bacterium are M77832 (19), M87484 (11), and X99636 (6).  相似文献   

16.
The chemical structure of the lipid A of the lipopolysaccharide component isolated from Flavobacterium meningosepticum IFO 12535 was elucidated. Methylation and nuclear magnetic resonance analyses showed that two kinds of hydrophilic backbone exist in the free lipid A: a β (1→6)-linked 2-amino-2-deoxy-d-glucose, which is usually present in enterobacterial lipid A’s, and a 2-amino-6-O-(2,3-diamino-2,3-dideoxy-β-d-glucopyranosyl)-2-deoxy-d-glucose, in a molar ratio of 1.00:0.35. Both backbones were α-glycosidically phosphorylated in position 1, and the hydroxyl groups at positions 4, 4′, and 6′ were unsubstituted. Liquid secondary ion-mass spectrometry revealed a pseudomolecular ion at m/z 1673 [M-H] as a major monophosphoryl lipid A component carrying five acyl groups. Fatty acid analysis showed that the lipid A contained 1 mol each of amide-linked (R)-3-OH iC17:0, ester-linked (R)-3-OH iC15:0, amide-linked (R)-3-O-(iC15:0)-iC17:0, and both amide- and ester-linked (R)-3-OH C16:0. Fatty acid distribution analyses using several mass spectrometry determinations demonstrated that the former two constituents were distributed on positions 2 and 3 of the reducing terminal unit of the backbones and that the latter two were attached to the 2′ and 3′ positions in the nonreducing terminal residue.Lipopolysaccharide (LPS) is known to act as an endotoxin that mediates pathophysiological changes such as fever and shock which occur in the course of severe gram-negative bacterial infection (5, 18, 27). The pathophysiological activity of LPS depends on the chemical structure of the hydrophobic portion called lipid A, the biologically active center of LPS (12, 14), which generally consists of a β(1→6)-linked 2-amino-2-deoxy-d-glucose (GlcN) disaccharide carrying phosphate and fatty acid residues; many fine structural variations are observed in different bacterial families (38).Since many of the LPSs from various gram-negative bacteria cause similar endotoxic effects despite differences in chemical composition and positions of substitution, the chemical structure required for the activity does not seem to be very strict. It has been reported, however, that several lipid A forms, isolated from the LPSs of Porphyromonas gingivalis (16), Rhodobacter sphaeroides (22, 26) and Rhodobacter capsulatus (20), as well as chemically synthesized lipid A analogs (6, 12), which are structurally similar to the active-type lipid A, exhibit dramatically low endotoxicity. Biologically active lipid A has been found to be changed to completely nontoxic derivatives by simple chemical modifications (28, 29). These findings indicate that the biological activity of lipid A is controlled by the fine structural variations. The nontoxic or low-toxicity lipid A preparations are very important for the determination of the relationship between the chemical structure and biological activity of lipid A, and also for the systematic development of LPS antagonists. However, the essential structural requirements for the complete activity or nontoxicity of lipid A are still uncertain. It is, therefore, meaningful to study the chemical and biological properties of naturally occurring lipid A’s which possess a unique structure.Flavobacterium meningosepticum is an aerobic gram-negative rod which is known to cause meningitis and septicemia in newborn infants (4, 21). Interestingly, the bacterium does not induce Limulus gelation activity when tested with whole cells (32), strongly suggesting that the LPS is of low toxicity or nontoxic and that the lipid A must have a unique structure relative to other enterobacterial lipid A’s.In the present study, the chemical structure of lipid A isolated from F. meningosepticum LPS was characterized by compositional study, methylation analysis, mass spectrometry (MS), and nuclear magnetic resonance (NMR) spectroscopy.  相似文献   

17.
18.
α-l-Arabinofuranosidases I and II were purified from the culture filtrate of Aspergillus awamori IFO 4033 and had molecular weights of 81,000 and 62,000 and pIs of 3.3 and 3.6, respectively. Both enzymes had an optimum pH of 4.0 and an optimum temperature of 60°C and exhibited stability at pH values from 3 to 7 and at temperatures up to 60°C. The enzymes released arabinose from p-nitrophenyl-α-l-arabinofuranoside, O-α-l-arabinofuranosyl-(1→3)-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose, and arabinose-containing polysaccharides but not from O-β-d-xylopyranosyl-(1→2)-O-α-l-arabinofuranosyl-(1→3)-O-β-d-xylopyranosyl-(1→4)-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose. α-l-Arabinofuranosidase I also released arabinose from O-β-d-xylopy-ranosyl-(1→4)-[O-α-l-arabinofuranosyl-(1→3)]-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose. However, α-l-arabinofuranosidase II did not readily catalyze this hydrolysis reaction. α-l-Arabinofuranosidase I hydrolyzed all linkages that can occur between two α-l-arabinofuranosyl residues in the following order: (1→5) linkage > (1→3) linkage > (1→2) linkage. α-l-Arabinofuranosidase II hydrolyzed the linkages in the following order: (1→5) linkage > (1→2) linkage > (1→3) linkage. α-l-Arabinofuranosidase I preferentially hydrolyzed the (1→5) linkage of branched arabinotrisaccharide. On the other hand, α-l-arabinofuranosidase II preferentially hydrolyzed the (1→3) linkage in the same substrate. α-l-Arabinofuranosidase I released arabinose from the nonreducing terminus of arabinan, whereas α-l-arabinofuranosidase II preferentially hydrolyzed the arabinosyl side chain linkage of arabinan.Recently, it has been proven that l-arabinose selectively inhibits intestinal sucrase in a noncompetitive manner and reduces the glycemic response after sucrose ingestion in animals (33). Based on this observation, l-arabinose can be used as a physiologically functional sugar that inhibits sucrose digestion. Effective l-arabinose production is therefore important in the food industry. l-Arabinosyl residues are widely distributed in hemicelluloses, such as arabinan, arabinoxylan, gum arabic, and arabinogalactan, and the α-l-arabinofuranosidases (α-l-AFases) (EC 3.2.1.55) have proven to be essential tools for enzymatic degradation of hemicelluloses and structural studies of these compounds.α-l-AFases have been classified into two families of glycanases (families 51 and 54) on the basis of amino acid sequence similarities (11). The two families of α-l-AFases also differ in substrate specificity for arabinose-containing polysaccharides. Beldman et al. summarized the α-l-AFase classification based on substrate specificities (3). One group contains the Arafur A (family 51) enzymes, which exhibit very little or no activity with arabinose-containing polysaccharides. The other group contains the Arafur B (family 54) enzymes, which cleave arabinosyl side chains from polymers. However, this classification is too broad to define the substrate specificities of α-l-AFases. There have been many studies of the α-l-AFases (3, 12), especially the α-l-AFases of Aspergillus species (28, 1215, 17, 22, 23, 2832, 3639, 4143, 46). However, there have been only a few studies of the precise specificities of these α-l-AFases. In previous work, we elucidated the substrate specificities of α-l-AFases from Aspergillus niger 5-16 (17) and Bacillus subtilis 3-6 (16, 18), which should be classified in the Arafur A group and exhibit activity with arabinoxylooligosaccharides, synthetic methyl 2-O-, 3-O-, and 5-O-arabinofuranosyl-α-l-arabinofuranosides (arabinofuranobiosides) (20), and methyl 3,5-di-O-α-l-arabinofuranosyl-α-l-arabinofuranoside (arabinofuranotrioside) (19).In the present work, we purified two α-l-AFases from a culture filtrate of Aspergillus awamori IFO 4033 and determined the substrate specificities of these α-l-AFases by using arabinose-containing polysaccharides and the core oligosaccharides of arabinoxylan and arabinan.  相似文献   

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

20.
Alginate is a polysaccharide composed of β-d-mannuronic acid (M) and α-l-guluronic acid (G). An Azotobacter vinelandii alginate lyase gene, algL, was cloned, sequenced, and expressed in Escherichia coli. The deduced molecular mass of the corresponding protein is 41.4 kDa, but a signal peptide is cleaved off, leaving a mature protein of 39 kDa. Sixty-three percent of the amino acids in this mature protein are identical to those in AlgL from Pseudomonas aeruginosa. AlgL was partially purified, and the activity was found to be optimal at a pH of 8.1 to 8.4 and at 0.35 M NaCl. Divalent cations are not necessary for activity. The pI of the enzyme is 5.1. When an alginate rich in mannuronic acid was used as the substrate, the Km was found to be 4.6 × 10−4 M (sugar residues). AlgL was found to cleave M-M and M-G bonds but not G-M or G-G bonds. Bonds involving acetylated residues were also cleaved, but this activity may be sensitive to the extent of acetylation.

Alginate is a family of 1-4-linked copolymers of β-d-mannuronic acid (M) and α-l-guluronic acid (G). It is produced by brown algae and by some bacteria belonging to the genera Azotobacter and Pseudomonas (8, 17, 18, 31). The polymer is widely used in industry and biotechnology (36, 44), and the genetics of its biosynthesis in Pseudomonas aeruginosa has been extensively studied due to its role in the disease cystic fibrosis (33). In bacterial alginates, some of the M residues may be O-2- and/or O-3-acetylated (42). The polymer is initially synthesized as mannuronan, and the G residues are introduced at the polymer level by mannuronan C-5-epimerases (13, 22, 23). The epimerized alginates contain a mixture of blocks of consecutive G residues (G blocks), consecutive M residues (M blocks), and alternating M and G residues (MG blocks). Alginates from Pseudomonas sp. do not contain G blocks (42).Alginate lyases catalyze the depolymerization of alginates by β-elimination, generating a molecule containing 4-deoxy-l-erythro-hex-4-enepyranosyluronate at the nonreducing end. Such lyases have been found in organisms using alginate as a carbon source, in bacteriophages specific for alginate-producing organisms, and in alginate-producing bacteria (45). An alginate molecule may contain four different glycosidic bonds, M-M, G-M, M-G, or G-G, and the relative rates at which each of these bonds are cleaved vary among different lyases (36a). The lyases also differ in the extent to which they are affected by acetylation (35, 43, 46).Davidson et al. (10) described an Azotobacter vinelandii lyase which preferred M blocks as a substrate. Kennedy et al. (28) later reported the purification of periplasmic alginate lyases from A. vinelandii and from Azotobacter chroococcum which also seemed to prefer deacetylated, M-rich alginate. The activities of these enzymes were found to be optimal at pH 6.8 and 7.2, respectively, while the enzyme reported by Davidson et al. (10) was found to display optimal activity at pH 7.8.A gene, algL, encoding an alginate lyase has been cloned from P. aeruginosa (2, 41). The gene was found to be located in a cluster containing most of the genes necessary for the biosynthesis of alginate. A homologous gene cluster has recently been identified in A. vinelandii (38) and shown to encode an alginate lyase (32). In our previous report, we showed that plasmid pHE102, which contains a part of this gene cluster, contains a DNA sequence sharing homology with algL from P. aeruginosa (38). We have now subcloned, sequenced, and expressed this gene in Escherichia coli. The lyase was shown to preferentially cleave deacetylated M-M and M-G bonds, but acetylated substrates were also cleaved.  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