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 15 毫秒
1.
Pyruvate (Pyr) and α-ketoglutarate (αKg) accumulated when cells of Pseudomonas fluorescens NCIMB 11764 were cultivated on growth-limiting amounts of ammonia or cyanide and were shown to be responsible for the nonenzymatic removal of cyanide from culture fluids as previously reported (J.-L. Chen and D. A. Kunz, FEMS Microbiol. Lett. 156:61–67, 1997). The accumulation of keto acids in the medium paralleled the increase in cyanide-removing activity, with maximal activity (760 μmol of cyanide removed min−1 ml of culture fluid−1) being recovered after 72 h of cultivation, at which time the keto acid concentration was 23 mM. The reaction products that formed between the biologically formed keto acids and cyanide were unambiguously identified as the corresponding cyanohydrins by 13C nuclear magnetic resonance spectroscopy. Both the Pyr and α-Kg cyanohydrins were further metabolized by cell extracts and served also as nitrogenous growth substrates. Radiotracer experiments showed that CO2 (and NH3) were formed as enzymatic conversion products, with the keto acid being regenerated as a coproduct. Evidence that the enzyme responsible for cyanohydrin conversion is cyanide oxygenase, which was shown previously to be required for cyanide utilization, is based on results showing that (i) conversion occurred only when extracts were induced for the enzyme, (ii) conversion was oxygen and reduced-pyridine nucleotide dependent, and (iii) a mutant strain defective in the enzyme was unable to grow when it was provided with the cyanohydrins as a growth substrate. Pyr and αKg were further shown to protect cells from cyanide poisoning, and excretion of the two was directly linked to utilization of cyanide as a growth substrate. The results provide the basis for a new mechanism of cyanide detoxification and assimilation in which keto acids play an essential role.Cyanide is a notorious poison. Its inhibitory effect on respiration has been known since the 1920s, when Warburg and Keilin first demonstrated that it combines with trivalent iron in cytochrome oxidase (38, 40, 44). Although highly toxic, it is a normal part of our environment for which mechanisms of biological formation (cyanogenesis) and detoxification exist (8, 22, 42). Cyanide also arises from various industrial practices such as steel coking, electroplating, and mining, but significant accumulations in the environment probably do not occur because of its highly reactive nature (13, 18, 41, 46). The interactions between microorganisms and cyanide, however, remain of interest, since the mechanisms of tolerance and assimilation are poorly understood. A number of reports documenting the ability of microorganisms to grow on cyanide have appeared, but the biochemical basis of these abilities has remained largely obscure. Most studies have reported its ability to serve as a nitrogen source only, since at the concentrations needed for it to serve as a carbon source, it is too toxic (15, 24). As far as is known, growth on cyanide requires that it be enzymatically converted to ammonia. Once formed it can then be readily incorporated into cellular macromolecules by established mechanisms (31). Two separate conversions have been described. They are hydrolytic and oxidative conversion, and they yield formic acid and carbon dioxide as reaction by-products, respectively. The enzyme responsible for hydrolytic conversion has variously been described as cyanidase, cyanide dihydratase, or cyanide nitrilase (CNN), and it catalyzes the reaction shown in equation 1. 1 Mechanistically, CNNs resemble other nitrilases (e.g., EC 3.5.5.1) that catalyze the direct conversion of organic nitriles into an organic acid and ammonia but for which the substrate range appears to be limited to cyanide. The involvement of CNNs in cyanide metabolism has been reported for Alcaligenes xylosooxidans subsp. denitrificans (19, 20), Bacillus pumilus (30), and Pseudomonas sp. (45). Oxidative conversion is mediated by an enzyme described as cyanide oxygenase (CNO). This enzyme has been described for Pseudomonas fluorescens NCIMB 11764 only (15, 2326). Recent work in our laboratory has shown that CNO functions as a monooxygenase, since a single atom of molecular oxygen was shown to be incorporated during conversion (43). Since the other atom of oxygen in CO2 was shown to be derived from water, a reaction mechanism in which cyanide undergoes initial monooxygenative attack to give an unknown intermediate (X-OH) as shown in equation 2 was proposed (43). 2 Further hydrolysis of X-OH is then expected to give CO2 and NH3 as shown in equation 3. 3 The nature of X-OH and whether an additional enzyme is required for its conversion are unknown. Interestingly, NCIMB 11764 also elaborates a CNN, but only CNO has been shown to be physiologically required for cyanide utilization (26). This conclusion was reached after it was found that mutants unable to grow on cyanide did not make CNO but could still elaborate CNN.CNO is induced when cyanide (KCN) is added to nitrogen-limited cells (4, 26). This approach for obtaining cells induced for the enzyme is more convenient than growing cells on cyanide, which requires several days of fed-batch cultivation. During the course of experiments aimed at optimizing CNO induction, we discovered that the consumption of cyanide and the appearance of CNO activity in cell extracts were not concomitant (3, 4). Further experiments showed that cyanide consumption independent of that catalyzed by CNO occurred nonenzymatically and that a reaction between cyanide and a metabolite excreted into the medium was responsible for cyanide’s removal (4). Since cyanide-removing activity in culture fluids consistently copurified with iron-chelating activity, it was concluded that the responsible metabolite was a siderophore, but further identification of this siderophore was not achieved. Here we report that the compounds responsible for nonenzymatic cyanide removal are α-keto acids, namely, pyruvate (Pyr) and α-ketoglutarate (αKg). These findings help explain the earlier reported involvement of a putative siderophore, since these compounds can act as iron chelators (10, 35). However, the additional ability to serve also as effective cyanide-scavenging agents has not generally been recognized. Both Pyr and αKg were excreted into the medium when P. fluorescens NCIMB 11764 was grown on nitrogen-limiting amounts of ammonia or cyanide as a nitrogen source, and we now demonstrate that these metabolites play an essential role in the utilization of cyanide as a growth substrate.  相似文献   

2.
Prostaglandin biosynthesis was studied in the rat uterus during the oestrous cycle. Uterine homogenates were incubated for 20 minutes in the presence of exogenous substrate (2.10−5M). PGF and PGE2 were measured by R.I.A.. A sharp peak PGF and a smaller peak of PGE2 were observed at prooestrus, 20 h. Another small PGE2 peak occurred at dioestrus II, 15 h. The lowest values of both PGs were found on dioestrus, 15 h. Plasma oestradiol concentration were highest at proestrus, 15 h and 20 h. A sharp progesterone peak occurred at prooestrus, 20 h. The PGF peak is next to the oestradiol peak and is superimposable or lags slightly beyond the progesterone peak.Incubation with 14C arachidonic acid and subsequent analysis of extracts by TLC and scanning showed that the major metabolite is PGI2, identified as 6 keto PGF. The conversion rate of arachidonic acid into 6 keto PGF is 5 times higher than into PGF. 6 keto PGF was further identified by GC/MS. No significant difference was observed between 6 keto PGF production during oestrus and dioestrus.  相似文献   

3.
The recycling of 5-methylthioribose (MTR) to methionine in avocado (Persea americana Mill, cv Hass) and tomato (Lycopersicum esculentum Mill, cv unknown) was examined. [14CH3]MTR was not metabolized in cell free extract from avocado fruit. Either [14CH3]MTR plus ATP or [14CH3]5-methylthioribose-1-phosphate (MTR-1-P) alone, however, were metabolized to two new products by these extracts. MTR kinase activity has previously been detected in these fruit extracts. These data indicate that MTR must be converted to MTR-1-P by MTR kinase before further metabolism can occur. The products of MTR-1-P metabolism were tentatively identified as α-keto-γ-methylthiobutyric acid (α-KMB) and α-hydroxy-γ-methylthiobutyric acid (α-HMB) by chromatography in several solvent systems. [35S]α-KMB was found to be further metabolized to methionine and α-HMB by these extracts, whereas α-HMB was not. However, α-HMB inhibited the conversion of α-KMB to methionine. Both [U-14C]α-KMB and [U-14C]methionine, but not [U-14C]α-HMB, were converted to ethylene in tomato pericarp tissue. In addition, aminoethoxyvinylglycine inhibited the conversion of α-KMB to ethylene. These data suggest that the recycling pathway leading to ethylene is MTR → MTR-1-P → α-KMB → methionine → S-adenosylmethionine → 1-aminocyclopropane-1-carboxylic acid → ethylene.  相似文献   

4.
l-Hydroxyproline (4-hydroxyproline) mainly exists in collagen, and most bacteria cannot metabolize this hydroxyamino acid. Pseudomonas putida and Pseudomonas aeruginosa convert l-hydroxyproline to α-ketoglutarate via four hypothetical enzymatic steps different from known mammalian pathways, but the molecular background is rather unclear. Here, we identified and characterized for the first time two novel enzymes, d-hydroxyproline dehydrogenase and Δ1-pyrroline-4-hydroxy-2-carboxylate (Pyr4H2C) deaminase, involved in this hypothetical pathway. These genes were clustered together with genes encoding other catalytic enzymes on the bacterial genomes. d-Hydroxyproline dehydrogenases from P. putida and P. aeruginosa were completely different from known bacterial proline dehydrogenases and showed similar high specificity for substrate (d-hydroxyproline) and some artificial electron acceptor(s). On the other hand, the former is a homomeric enzyme only containing FAD as a prosthetic group, whereas the latter is a novel heterododecameric structure consisting of three different subunits (α4β4γ4), and two FADs, FMN, and [2Fe-2S] iron-sulfur cluster were contained in αβγ of the heterotrimeric unit. These results suggested that the l-hydroxyproline pathway clearly evolved convergently in P. putida and P. aeruginosa. Pyr4H2C deaminase is a unique member of the dihydrodipicolinate synthase/N-acetylneuraminate lyase protein family, and its activity was competitively inhibited by pyruvate, a common substrate for other dihydrodipicolinate synthase/N-acetylneuraminate lyase proteins. Furthermore, disruption of Pyr4H2C deaminase genes led to loss of growth on l-hydroxyproline (as well as d-hydroxyproline) but not l- and d-proline, indicating that this pathway is related only to l-hydroxyproline degradation, which is not linked to proline metabolism.  相似文献   

5.
A pathogenic V67M mutation occurs at the E11 helical position within the heme pockets of variant human fetal and adult hemoglobins (Hb). Subsequent post-translational modification of Met to Asp was reported in γ subunits of human fetal Hb Toms River (γ67(E11)Val → Met) and β subunits of adult Hb (HbA) Bristol-Alesha (β67(E11)Val → Met) that were associated with hemolytic anemia. Using kinetic, proteomic, and crystal structural analysis, we were able to show that the Met → Asp transformation involves heme cycling through its oxoferryl state in the recombinant versions of both proteins. The conversion to Met and Asp enhanced the spontaneous autoxidation of the mutants relative to wild-type HbA and human fetal Hb, and the levels of Asp were elevated with increasing levels of hydrogen peroxide (H2O2). Using H218O2, we verified incorporation of 18O into the Asp carboxyl side chain confirming the role of H2O2 in the oxidation of the Met side chain. Under similar experimental conditions, there was no conversion to Asp at the αMet(E11) position in the corresponding HbA Evans (α62(E11)Val → Met). The crystal structures of the three recombinant Met(E11) mutants revealed similar thioether side chain orientations. However, as in the solution experiments, autoxidation of the Hb mutant crystals leads to electron density maps indicative of Asp(E11) formation in β subunits but not in α subunits. This novel post-translational modification highlights the nonequivalence of human Hb α, β, and γ subunits with respect to redox reactivity and may have direct implications to α/β hemoglobinopathies and design of oxidatively stable Hb-based oxygen therapeutics.  相似文献   

6.
Two Salmonella typhimurium strains, which could be used as sources for the leucine biosynthetic intermediates α- and β-isopropylmalate were constructed by a series of P22-mediated transductions. One strain, JK527 [flr-19 leuA2010 Δ(leuD-ara)798 fol-162], accumulated and excreted α-isopropylmalate, whereas the second strain, JK553 (flr-19 leuA2010 leuB698), accumulated and excreted α- and β-isopropylmalate. The yield of α-isopropylmalate isolated from the culture medium of JK527 was more than five times the amount obtained from a comparable volume of medium in which Neurospora crassa strain FLR92-1-216 (normally used as the source for α- and β-isopropylmalate) was grown. Not only was the yield greater, but S. typhimurium strains are much easier to handle and grow to saturation much faster than N. crassa strains. The combination of the two regulatory mutations flr-19, which results in constitutive expression of the leucine operon, and leuA2010, which renders the first leucine-specific biosynthetic enzyme insensitive to feedback inhibition by leucine, generated limitations in the production of valine and pantothenic acid. The efficient, irreversible, and unregulated conversion of α-ketoisovaleric acid into α-isopropylmalate (α-isopropylmalate synthetase Km for α-ketoisovaleric acid, 6 × 10−5 M) severely restricted the amount of α-ketoisovaleric acid available for conversion into valine and pantothenic acid (ketopantoate hydroxymethyltransferase Km for α-ketoisovaleric acid, 1.1 × 10−3 M; transaminase B Km for α-ketoisovaleric acid, 2 × 10−3 M).  相似文献   

7.
The epithelial sodium channel is a multimeric protein formed by three homologous subunits: α, β, and γ; each subunit contains only two transmembrane domains. The level of expression of each of the subunits is markedly different in various Na+ absorbing epithelia raising the possibility that channels with different subunit composition can function in vivo. We have examined the functional properties of channels formed by the association of α with β and of α with γ in the Xenopus oocyte expression system using two-microelectrode voltage clamp and patch-clamp techniques. We found that αβ channels differ from αγ channels in the following functional properties: (a) αβ channels expressed larger Na+ than Li+ currents (INa+/ILi+ 1.2) whereas αγ channels expressed smaller Na+ than Li+ currents (INa+/ILi+ 0.55); (b) the Michaelis Menten constants (K m) of activation of current by increasing concentrations of external Na+ and Li+ of αβ channels were larger (K m > 180 mM) than those of αγ channels (K m of 35 and 50 mM, respectively); (c) single channel conductances of αβ channels (5.1 pS for Na+ and 4.2 pS for Li+) were smaller than those of αγ channels (6.5 pS for Na+ and 10.8 pS for Li+); (d) the half-inhibition constant (K i) of amiloride was 20-fold larger for αβ channels than for αγ channels whereas the K i of guanidinium was equal for both αβ and αγ. To identify the domains in the channel subunits involved in amiloride binding, we constructed several chimeras that contained the amino terminus of the γ subunit and the carboxy terminus of the β subunit. A stretch of 15 amino acids, immediately before the second transmembrane domain of the β subunit, was identified as the domain conferring lower amiloride affinity to the αβ channels. We provide evidence for the existence of two distinct binding sites for the amiloride molecule: one for the guanidium moiety and another for the pyrazine ring. At least two subunits α with β or γ contribute to these binding sites. Finally, we show that the most likely stoichiometry of αβ and αγ channels is 1α:1β and 1α:1γ, respectively.  相似文献   

8.
Lipid-linked oligosaccharides were synthesized with the particulate enzyme preparation from mung bean (Phaseolus aureus) seedlings in the presence of GDP-[14C] mannose. The oligosaccharides were released from the lipids by mild acid hydrolysis and purified by several passages on Biogel P-4 columns. Five different oligosaccharides were purified in this way. Based on their relative elution constants (Kd) compared to a variety of standard oligosaccharides, they were sized as (mannose-acetylglucosamine) Man7GlcNAc2, Man5GlcNAc2, Man3GlcNAc2, Man2GlcNAc2, and ManGlcNAc2. These oligosaccharides were treated with endoglucosaminidase H and α- and β-mannosidase, and the products were examined on Biogel P-4 columns. They also were subjected to a number of chemical treatments including analysis of the reducing sugar by NaB3H4 reduction, methylation analysis, and in some cases acetolysis. From these data, the likely structures of these oligosaccharides are as follows: E, Manβ-GlcNAc-GlcNAc; D, Manα1→3Manβ-GlcNAc-GlcNAc; C, Manα1→2Manα1→3Manβ-GlcNAc-GlcNAc; B, Manα1→2Manα1→2Manα1→ 3(Manα1→6)Manβ-GlcNAc-GlcNAc; and A, Manα1→2Manα1→ 2Manα1→3(Manα1→ [Manα1→6]Manα1→6) Manβ-GlcNAc-GlcNAc. The synthesis of the Man7GlcNAc2 was greatly diminished when tunicamycin (10 μg/ml) was added to the incubation mixtures.  相似文献   

9.
An R-phycoerythrin (R-PE) was isolated by gel filtrations on Sepharose CL-4B and Sephadex G-150 from the phycobiliprotein extract of the marine red macroalga Polysiphonia urceolata Grev and further purified by ion exchange chromatography on DEAE-Sepharose Fast Flow. The purified R-PE showed three absorption peaks at 498 nm, 538 nm, 566 nm and one fluorescent emission maximum at 577 nm. Although the R-PE showed a single band on the examination by native PAGE, it exhibited two very close bands at pH about 4.7 in native isoelectric focusing (IEF). Polypeptide analysis of the R-PE demonstrated that it contained four chromophore-carrying subunits, α18.2, β20.6, γ31.6 (γ''), γ34.6 (γ), and no colorless polypeptide; its subunit composition was 6α18.2:6β20.6:1 γ31.6:2γ34.6. The α and β subunits were distributed within a acidic pH range from 5.0 to 6.0 in denaturing IEF and the γ subunits were in a basic pH range from 7.6 to 8.1. These results reveal that the prepared R-PE may exist in two hexamers of γ (αβ)3 γ (αβ)3γ'' and γ (αβ)3 γ''(αβ)3 γ and that the R-PE participate in the rod domain assembly of P. urceolata phycobilisomes by stacking each of its trimer (αβ)3 face-to-face with the aid of one γ subunit (γ or γ'').  相似文献   

10.
The transmission of infectious prions into different host species requires compatible prion protein (PrP) primary structures, and even one heterologous residue at a pivotal position can block prion infection. Mapping the key amino acid positions that govern cross-species prion conversion has not yet been possible, although certain residue positions have been identified as restrictive, including residues in the β22 loop region of PrP. To further define how β22 residues impact conversion, we investigated residue substitutions in PrPC using an in vitro prion conversion assay. Within the β22 loop, a tyrosine residue at position 169 is strictly conserved among mammals, and transgenic mice expressing mouse PrP having the Y169G, S170N, and N174T substitutions resist prion infection. To better understand the structural requirements of specific residues for conversion initiated by mouse prions, we substituted a diverse array of amino acids at position 169 of PrP. We found that the substitution of glycine, leucine, or glutamine at position 169 reduced conversion by ∼75%. In contrast, replacing tyrosine 169 with either of the bulky, aromatic residues, phenylalanine or tryptophan, supported efficient prion conversion. We propose a model based on a requirement for tightly interdigitating complementary amino acid side chains within specific domains of adjacent PrP molecules, known as “steric zippers,” to explain these results. Collectively, these studies suggest that an aromatic residue at position 169 supports efficient prion conversion.  相似文献   

11.
Central adiposity is a significant determinant of obesity-related hypertension risk, which may arise due to the pathogenic inflammatory nature of the abdominal fat depot. However, the influence of pro-inflammatory adipokines on blood pressure in the obese hypertensive phenotype has not been well established in Saudi subjects. As such, our study investigated whether inflammatory factors may represent useful biomarkers to delineate hypertension risk in a Saudi cohort with and without hypertension and/or diabetes mellitus type 2 (DMT2). Subjects were subdivided into four groups: healthy lean controls (age: 47.9±5.1 yr; BMI: 22.9±2.1 Kg/m2), non-hypertensive obese (age: 46.1±5.0 yr; BMI: 33.7±4.2 Kg/m2), hypertensive obese (age: 48.6±6.1 yr; BMI: 36.5±7.7 Kg/m2) and hypertensive obese with DMT2 (age: 50.8±6.0 yr; BMI: 35.3±6.7 Kg/m2). Anthropometric data were collected from all subjects and fasting blood samples were utilized for biochemical analysis. Serum angiotensin II (ANG II) levels were elevated in hypertensive obese (p<0.05) and hypertensive obese with DMT2 (p<0.001) compared with normotensive controls. Systolic blood pressure was positively associated with BMI (p<0.001), glucose (p<0.001), insulin (p<0.05), HOMA-IR (p<0.001), leptin (p<0.01), TNF-α (p<0.001) and ANG II (p<0.05). Associations between ANG II and TNF-α with systolic blood pressure remained significant after controlling for BMI. Additionally CRP (p<0.05), leptin (p<0.001) and leptin/adiponectin ratio (p<0.001) were also significantly associated with the hypertension phenotype. In conclusion our data suggests that circulating pro-inflammatory adipokines, particularly ANG II and, TNF-α, represent important factors associated with a hypertension phenotype and may directly contribute to predicting and exacerbating hypertension risk.  相似文献   

12.
The neural cell adhesion molecule L1 has been shown to function as a homophilic ligand in a variety of dynamic neurological processes. Here we demonstrate that the sixth immunoglobulin-like domain of human L1 (L1-Ig6) can function as a heterophilic ligand for multiple members of the integrin superfamily including αvβ3, αvβ1, α5β1, and αIIbβ3. The interaction between L1-Ig6 and αIIbβ3 was found to support the rapid attachment of activated human platelets, whereas a corresponding interaction with αvβ3 and αvβ1 supported the adhesion of umbilical vein endothelial cells. Mutation of the single Arg-Gly-Asp (RGD) motif in human L1-Ig6 effectively abrogated binding by the aforementioned integrins. A L1 peptide containing this RGD motif and corresponding flanking amino acids (PSITWRGDGRDLQEL) effectively blocked L1 integrin interactions and, as an immobilized ligand, supported adhesion via αvβ3, αvβ1, α5β1, and αIIbβ3. Whereas β3 integrin binding to L1-Ig6 was evident in the presence of either Ca2+, Mg2+, or Mn2+, a corresponding interaction with the β1 integrins was only observed in the presence of Mn2+. Furthermore, such Mn2+-dependent binding by α5β1 and αvβ1 was significantly inhibited by exogenous Ca2+. Our findings suggest that physiological levels of calcium will impose a hierarchy of integrin binding to L1 such that αvβ3 or active αIIbβ3 > αvβ1 > α5β1. Given that L1 can interact with multiple vascular or platelet integrins it is significant that we also present evidence for de novo L1 expression on blood vessels associated with certain neoplastic or inflammatory diseases. Together these findings suggest an expanded and novel role for L1 in vascular and thrombogenic processes.  相似文献   

13.
The marine diatom Amphora coffeaeformis required Ca2+ and bicarbonate for motility. Movement was inhibited by the Ca2+-blocking agents ruthenium red and α-isopropyl-α-[(N-methyl-N-homoveratryl)-α- aminopropyl]-3,4,5-trimethoxy phenyl acetonitrile and the metabolic energy uncoupler, carbonyl cyanide 3-chlorophenylhydrazone. 3-(3′,4-Dichlorophenyl)-1,1-Dimethyl urea was without effect on cells at a concentration that prevented O2 production in the light. Although Sr2+ could replace Ca2+ in the attachment of cells to glass, it did not substitute for Ca2+ in motility.  相似文献   

14.
The α1A-AR is thought to couple predominantly to the Gαq/PLC pathway and lead to phosphoinositide hydrolysis and calcium mobilization, although certain agonists acting at this receptor have been reported to trigger activation of arachidonic acid formation and MAPK pathways. For several G protein-coupled receptors (GPCRs) agonists can manifest a bias for activation of particular effector signaling output, i.e. not all agonists of a given GPCR generate responses through utilization of the same signaling cascade(s). Previous work with Gαq coupling-defective variants of α1A-AR, as well as a combination of Ca2+ channel blockers, uncovered cross-talk between α1A-AR and β2-AR that leads to potentiation of a Gαq-independent signaling cascade in response to α1A-AR activation. We hypothesized that molecules exist that act as biased agonists to selectively activate this pathway. In this report, isoproterenol (Iso), typically viewed as β-AR-selective agonist, was examined with respect to activation of α1A-AR. α1A-AR selective antagonists were used to specifically block Iso evoked signaling in different cellular backgrounds and confirm its action at α1A-AR. Iso induced signaling at α1A-AR was further interrogated by probing steps along the Gαq /PLC, Gαs and MAPK/ERK pathways. In HEK-293/EBNA cells transiently transduced with α1A-AR, and CHO_α1A-AR stable cells, Iso evoked low potency ERK activity as well as Ca2+ mobilization that could be blocked by α1A-AR selective antagonists. The kinetics of Iso induced Ca2+ transients differed from typical Gαq- mediated Ca2+ mobilization, lacking both the fast IP3R mediated response and the sustained phase of Ca2+ re-entry. Moreover, no inositol phosphate (IP) accumulation could be detected in either cell line after stimulation with Iso, but activation was accompanied by receptor internalization. Data are presented that indicate that Iso represents a novel type of α1A-AR partial agonist with signaling bias toward MAPK/ERK signaling cascade that is likely independent of coupling to Gαq.  相似文献   

15.
The integrin α4β7 plays an important role in lymphocyte homing to mucosal lymphoid tissues and has been shown to define a subpopulation of memory T cells capable of homing to intestinal sites. Here we have used a well-characterized intestinal virus, murine rotavirus, to investigate whether memory/effector function for an intestinal pathogen is associated with α4β7 expression. α4β7hi memory phenotype (CD44hi), α4β7 memory phenotype, and presumptively naive (CD44lo) CD8+ T lymphocytes from rotavirus-infected mice were sorted and transferred into Rag-2 (T- and B-cell-deficient) recipients that were chronically infected with murine rotavirus. α4β7hi memory phenotype CD8+ cells were highly efficient at clearing rotavirus infection, α4β7 memory cells were inefficient or ineffective, depending on the cell numbers transferred, and CD44lo cells were completely unable to clear chronic rotavirus infection. These data demonstrate that functional memory for rotavirus resides primarily in memory phenotype cells that display the mucosal homing receptor α4β7.  相似文献   

16.
During anaerobic growth of Klebsiella pneumoniae on citrate, 9.4 mmol of H2/mol of citrate (4-kPa partial pressure) was formed at the end of growth besides acetate, formate, and CO2. Upon addition of NiCl2 (36 μM) to the growth medium, hydrogen formation increased about 36% to 14.8 mmol/mol of citrate (6 kPa), and the cell yield increased about 15%. Cells that had been harvested and washed under anoxic conditions exhibited an H2-dependent formation of NAD(P)H in vivo. The reduction of internal NAD(P)+ was also achieved by the addition of formate. In crude extracts, the H2:NAD+ oxidoreductase activity was 0.13 μmol min−1 mg−1, and 76% of this activity was found in the washed membrane fraction. The highest specific activities of the membrane fraction were observed in 50 mM potassium phosphate, with 1.6 μmol of NADPH formed min−1 mg−1 at pH 7.0 and 1.7 μmol of NADH formed min−1 mg−1 at pH 9.5. In the presence of the protonophore carbonyl cyanide m-chlorophenylhydrazone and the Na+/H+ antiporter monensin, the H2-dependent reduction of NAD+ by membrane vesicles decreased only slightly (about 16%). The NADP+- or NAD+-reducing hydrogenases were solubilized from the membranes with the detergent lauryldimethylamine-N-oxide or Triton X-100. NAD(P)H formation with H2 as electron donor, therefore, does not depend on an energized state of the membrane. It is proposed that hydrogen which is formed by K. pneumoniae during citrate fermentation is recaptured by a novel membrane-bound, oxygen-sensitive H2:NAD(P)+ oxidoreductase that provides reducing equivalents for the synthesis of cell material.  相似文献   

17.
18.

Background

Hemostasis and thrombosis are regulated by agonist-induced activation of platelet integrin αIIbβ3. Integrin activation, in turn is mediated by cellular signaling via protein kinases and protein phosphatases. Although the catalytic subunit of protein phosphatase 1 (PP1c) interacts with αIIbβ3, the role of PP1c in platelet reactivity is unclear.

Methodology/Principal Findings

Using γ isoform of PP1c deficient mice (PP1cγ−/−), we show that the platelets have moderately decreased soluble fibrinogen binding and aggregation to low concentrations of thrombin or protease-activated receptor 4 (PAR4)-activating peptide but not to adenosine diphosphate (ADP), collagen or collagen-related peptide (CRP). Thrombin-stimulated PP1cγ−/− platelets showed decreased αIIbβ3 activation despite comparable levels of αIIbβ3, PAR3, PAR4 expression and normal granule secretion. Functions regulated by outside-in integrin αIIbβ3 signaling like adhesion to immobilized fibrinogen and clot retraction were not altered in PP1cγ−/− platelets. Thrombus formation induced by a light/dye injury in the cremaster muscle venules was significantly delayed in PP1cγ−/− mice. Phosphorylation of glycogen synthase kinase (GSK3)β-serine 9 that promotes platelet function, was reduced in thrombin-stimulated PP1cγ−/− platelets by an AKT independent mechanism. Inhibition of GSK3β partially abolished the difference in fibrinogen binding between thrombin-stimulated wild type and PP1cγ−/− platelets.

Conclusions/Significance

These studies illustrate a role for PP1cγ in maintaining GSK3β-serine9 phosphorylation downstream of thrombin signaling and promoting thrombus formation via fibrinogen binding and platelet aggregation.  相似文献   

19.
The interaction of redox-active copper ions with misfolded amyloid β (Aβ) is linked to production of reactive oxygen species (ROS), which has been associated with oxidative stress and neuronal damages in Alzheimer disease. Despite intensive studies, it is still not conclusive how the interaction of Cu+/Cu2+ with Aβ aggregates leads to ROS production even at the in vitro level. In this study, we examined the interaction between Cu+/Cu2+ and Aβ fibrils by solid-state NMR (SSNMR) and other spectroscopic methods. Our photometric studies confirmed the production of ∼60 μm hydrogen peroxide (H2O2) from a solution of 20 μm Cu2+ ions in complex with Aβ(1–40) in fibrils ([Cu2+]/[Aβ] = 0.4) within 2 h of incubation after addition of biological reducing agent ascorbate at the physiological concentration (∼1 mm). Furthermore, SSNMR 1H T1 measurements demonstrated that during ROS production the conversion of paramagnetic Cu2+ into diamagnetic Cu+ occurs while the reactive Cu+ ions remain bound to the amyloid fibrils. The results also suggest that O2 is required for rapid recycling of Cu+ bound to Aβ back to Cu2+, which allows for continuous production of H2O2. Both 13C and 15N SSNMR results show that Cu+ coordinates to Aβ(1–40) fibrils primarily through the side chain Nδ of both His-13 and His-14, suggesting major rearrangements from the Cu2+ coordination via Nϵ in the redox cycle. 13C SSNMR chemical shift analysis suggests that the overall Aβ conformations are largely unaffected by Cu+ binding. These results present crucial site-specific evidence of how the full-length Aβ in amyloid fibrils offers catalytic Cu+ centers.  相似文献   

20.
We have further characterized at the single channel level the properties of epithelial sodium channels formed by coexpression of α with either wild-type β or γ subunits and α with carboxy-terminal truncated β (βT) or γ (γT) subunits in Xenopus laevis oocytes. αβ and αβT channels (9.6 and 8.7 pS, respectively, with 150 mM Li+) were found to be constitutively open. Only upon inclusion of 1 μM amiloride in the pipette solution could channel activity be resolved; both channel types had short open and closed times. Mean channel open probability (P o) for αβ was 0.54 and for αβT was 0.50. In comparison, αγ and αγT channels exhibited different kinetics: αγ channels (6.7 pS in Li+) had either long open times with short closings, resulting in a high P o (0.78), or short openings with long closed times, resulting in a low P o (0.16). The mean P o for all αγ channels was 0.48. αγT (6.6 pS in Li+) behaved as a single population of channels with distinct kinetics: mean open time of 1.2 s and closed time of 0.4 s, with a mean P o of 0.6, similar to that of αγ. Inclusion of 0.1 μM amiloride in the pipette solution reduced the mean open time of αγT to 151 ms without significantly altering the closed time. We also examined the kinetics of amiloride block of αβ, αβT (1 μM amiloride), and αγT (0.1 μM amiloride) channels. αβ and αβT had similar blocking and unblocking rate constants, whereas the unblocking rate constant for αγT was 10-fold slower than αβT. Our results indicate that subunit composition of ENaC is a main determinant of P o. In addition, channel kinetics and P o are not altered by carboxy-terminal deletion in the β subunit, whereas a similar deletion in the γ subunit affects channel kinetics but not P o.  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