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 62 毫秒
1.
2.
In the recent past, we have observed a possible role of 10398A and 16189C mtDNA and PGC1α p.Thr394Thr (rs2970847) and p.Gly482Ser (rs8192673) variant genotypes providing susceptibility/protection against type 2 diabetes mellitus (T2DM) in two North Indian population groups. These initial observations encouraged us to look at the candidate genes in combination with –866G/A (rs659366) polymorphism in uncoupling protein 2 (UCP2) in a single study of a relatively large sample size, constituted of both the cohorts, to unravel an interesting outcome of an additive interaction in-between the studied genes. In a total of 1,686 individuals (762 cases and 924 controls) belonging to Indo-European linguistic group from North India, a comparison of risk genotype combinations of: UCP2–866GG, mtDNA 10398A and PGC1α p.Thr394Thr or p.Gly482Ser against the protective genotypes: UCP2–866XA, mtDNA 10398G and PGC1α p.Thr394Thr (nominal P value = 1.75 × 10−14, Odds ratio, OR = 5.29, 3.40–8.22 at 95% CI) or PGC1α p.Gly482Ser (nominal p value = 4.42 × 10−24, OR = 8.59, 5.53–13.35 at 95% CI), showed a highly significant difference and increased ORs. In a complex disease, it is always encouraging to find an additive interaction of multiple small effects of the studied candidate gene variations. An erratum to this article is available at .  相似文献   

3.
The influence of environmental pH on biological activity of Bifidobacterium longum CRL 849 grown in MRS-raffinose was evaluated. At pH 6.0, 5.5 and 5.0, raffinose was completely consumed by this microorganism, showing different consumption rates at each pH value (between 3.03 and 0.76 mmol l−1 h−1). At pH 4.5, the growth was lowest. The removal of raffinose was due to the α-galactosidase (α-gal) activity of this bifidobacteria, which was highest at pH 6.0–5.5 (1,280–1,223 mU ml−1). The production of β-glucosidase (β-glu) showed a similar pattern to α-gal activity with major values. The yield of organic acids produced during raffinose consumption was also highest at pH 6.0–5.5. The results of this study will allow the selection of the optimum growth conditions of B. longum CRL 849, with elevated levels of α-gal to be used in the reduction of nondigestible α-oligosaccharide in soy products and β-glu activities involved in isoflavone conversion to bioactive forms when used as starter culture.  相似文献   

4.
Highly precise and accurate measurements of very small NMR cross-correlated relaxation rates, namely those between protein HiN–Ni and Ci−1α–Ci−1′ dipoles, are demonstrated with an error of 0.03 s−1 for GB3. Because the projection angles between the two dipole vectors are very close to the magic angle the rates range only from −0.2 to +0.2 s−1. Small changes of the average vector orientations have a dramatic impact on the relative values. The rates suggest deviation from idealized peptide plane geometry caused by twists around the C′–N bonds and/or pyramidalization of the nitrogen atoms. A clear alternating pattern along the sequence is observed in β strands 1, 3 and 4 of GB3, where the side chains of almost all residues with large positive rates are solvent exposed. In the α helix all rates are relatively large and positive. Some of the currently most accurate structures of GB3 determined by both high resolution X-ray crystallography and NMR are in satisfactory agreement with the experimental rates in the helix and β strand 3, but not in the loops and the two central strands of the sheet for which no alternating pattern is predicted.  相似文献   

5.
The metal–thiolate connectivity of recombinant Cd7-MT10 metallothionein from the sea mussel Mytilus galloprovincialis has been investigated for the first time by means of multinuclear, multidimensional NMR spectroscopy. The internal backbone dynamics of the protein have been assessed by the analysis of 15N T 1 and T 2 relaxation times and steady state {1H}–15N heteronuclear NOEs. The 113Cd NMR spectrum of mussel MT10 shows unique features, with a remarkably wide dispersion (210 ppm) of 113Cd NMR signals. The complete assignment of cysteine Hα and Hβ proton resonances and the analysis of 2D 113Cd–113Cd COSY and 1H–113Cd HMQC type spectra allowed us to identify a four metal–thiolate cluster (α-domain) and a three metal–thiolate cluster (β-domain), located at the N-terminal and the C-terminal, respectively. With respect to vertebrate MTs, the mussel MT10 displays an inversion of the α and β domains inside the chain, similar to what observed in the echinoderm MT-A. Moreover, unlike the MTs characterized so far, the α-domain of mussel Cd7-MT10 is of the form M4S12 instead of M4S11, and has a novel topology. The β-domain has a metal–thiolate binding pattern similar to other vertebrate MTs, but it is conformationally more rigid. This feature is quite unusual for MTs, in which the β-domain displays a more disordered conformation than the α-domain. It is concluded that in mussel Cd7-MT10, the spacing of cysteine residues and the plasticity of the protein backbone (due to the high number of glycine residues) increase the adaptability of the protein backbone towards enfolding around the metal–thiolate clusters, resulting in minimal alterations of the ideal tetrahedral geometry around the metal centres.  相似文献   

6.
Summary A Monte Carlo simulation is proposed to study the dynamics of helper T-cells (N H) and viral (N V) populations in an immune response model relevant to HIV. Cellular states are binary variables and the interactions are described by logical expressions. Viral population shows a nonmonotonic growth before reaching a constant value while helper T-cells grow to a constant after a relaxation/reaction time. Initially, the population of helper cells grows with time with a power-law, N Ht β, before reaching the steady-state; the growth exponent β increases systematically (β ≈ 1 – 2) with the mutation rate (P mut≈0.1–0.4). The critical recovery time (t c) increases exponentially with the viral mutation, t cAe αP mut , with α=4.52±0.29 in low mutation regime and α=15.21±1.41 in high mutation regime. The equilibrium population of helper T-cell declines slowly with P mut and collapses at ∼ 0.40; the viral population exhibits a reverse trend, i.e., a slow increase before the burst around the same mutation regime.  相似文献   

7.
Killing of wild-type spores of Bacillus subtilis by t-butyl hydroperoxide, cumene hydroperoxide and peracetic acid was not through DNA damage, as shown by the absence of mutations in the survivors and the identical sensitivity of spores of strains with or without a recA mutation. In contrast, B. subtilis spores (termed αβ) lacking the DNA protective α/β-type small, acid-soluble spore proteins (SASP) were more sensitive to t-butyl hydroperoxide and cumene hydroperoxide, and their killing was in large part through DNA damage, as shown by the high frequency of mutations in the survivors and the greater sensitivity of αβ recA spores. Analysis of t-butyl hydroperoxide-treated spores showed that generation of DNA damage in αβ spores was more rapid than in wild-type spores; α/β-type SASP also protected against DNA strand breakage in vitro caused by t-butyl hydroperoxide. α/β-Type SASP appeared to play no role in protection of spores from killing by peracetic acid; wild-type and αβ spores exhibited identical peracetic acid sensitivity and their killing by this agent appeared to be not through DNA damage. Received 17 December 1996/ Accepted in revised form 13 March 1997  相似文献   

8.
Nitrile hydratases (NHase), which catalyze the hydration of nitriles to amides, have an unusual Fe3+ or Co3+ center with two modified Cys ligands: cysteine sulfininate (Cys-SO2 ) and either cysteine sulfenic acid or cysteine sulfenate [Cys-SO(H)]. Two catalytic mechanisms have been proposed. One is that the sulfenyl oxygen activates a water molecule, enabling nucleophilic attack on the nitrile carbon. The other is that the Ser ligand ionizes the strictly conserved Tyr, activating a water molecule. Here, we characterized mutants of Fe-type NHase from Rhodococcus erythropolis N771, replacing the Ser and Tyr residues, αS113A and βY72F. The αS113A mutation partially affected catalytic activity and did not change the pH profiles of the kinetic parameters. UV–vis absorption spectra indicated that the electronic state of the Fe center was altered by the αS113A mutation, but the changes could be prevented by a competitive inhibitor, n-butyric acid. The overall structure of the αS113A mutant was similar to that of the wild type, but significant changes were observed around the catalytic cavity. Like the UV–vis spectra, the changes were compensated by the substrate or product. The Ser ligand is important for the structure around the catalytic cavity, but is not essential for catalysis. The βY72F mutant exhibited no activity. The structure of the βY72F mutant was highly conserved but was found to be the inactivated state, with αCys114-SO(H) oxidized to Cys-SO2 , suggesting that βTyr72 affected the electronic state of the Fe center. The catalytic mechanism is discussed on the basis of the results obtained.  相似文献   

9.
A mannan of Candida glabrata IFO 0622 digested by Arthrobacter exo-α-mannosidase and a β-1,2-linked mannobiose obtained from the parent mannan by acid treatment was analyzed using 13C nuclear magnetic resonance spectroscopy. The results show that the β-1,2-linked mannobiosyl residue is esterified to a phosphate group through position C-1 in the α-configuration, Manβ1– 2Manα1–HPO3–. The results of immunochemical assays of these mannans using the commercial antigenic factor sera of the genus Candida (Candida Check, Iatron) indicate that the main recognition site of serum no. 6 in this kit is the mannotetraosyl side-chain Manβ1–2Manα1– 2Manα1–2Man in C. glabrata mannan and also suggest that the phosphate-containing unit (such as Manβ1– 2Manα1–HPO3– in this mannan) behaves as one of the antigenic determinants of serum no. 6, but not of serum no. 5. Therefore, the present and previous findings indicate that serum no. 5 recognizes relatively longer β-1,2-linked oligomannosyl side-chains, Manβ1–[2Manβ1–]n 2Man (n = 1–6), attached to the phosphate groups previously observed in the cell wall mannans of Candida albicans, Candida stellatoidea, and Candida tropicalis. Received: 18 March 1997 / Accepted: 16 September 1997  相似文献   

10.
Songbirds are widely studied to investigate the hormonal control of behavior. However, little is known about the effects of steroids on neurotransmission in these birds. We used electrophysiological and pharmacological techniques to characterize γ-aminobutyric acid (GABA) type A receptors (GABAA) of primary cultured telencephalic and hippocampal neurons from developing zebra finches. Additionally, their modulation by 17β-estradiol(E2), 5α- and 5β-dihydrotestosterone (DHT), 5α- and 5β-pregnan-3α-ol-20-one, and corticosterone was examined. Whole-cell GABA-evoked currents were inhibited by picrotoxin (10 μmol l−1) and bicuculline methiodide (10 μmol l−1) and potentiated by pentobarbital (100 μmol l−1) and propofol (3 μmol l−1). Loreclezole (10 μmol l−1) potentiated GABA-evoked currents, suggesting the presence of β2, β3 and/or β4 subunits. Diazepam (1 μmol l−1) potentiated currents, while Zn2+ (1 μmol l−1) caused no inhibition, indicating the presence of γ subunits. 5α- and 5β-Pregnan-3α-ol-20-one (100 nmol l−1) potentiated currents, whereas E2 (1 μmol l−1), 5α- and 5β-DHT (1 μmol l−1), and corticosterone (10 μmol l−1) had no detectable effect. We conclude that zebra finch telencephalic and hippocampal GABAA receptors include α, β, and γ subunits and are similar to their mammalian counterparts in both their biophysical and pharmacological properties. Additionally, GABA-evoked currents are greatly potentiated by 5α- and 5β-pregnan-3α-ol-20-one but show little or no acute modulation by sex steroids or corticosterone. Accepted: 12 November 1997  相似文献   

11.
Photosynthetic parameters of phytoplankton and sea ice algae from landfast sea ice of the Chukchi Sea off Point Barrow, Alaska, were assessed in spring 2005 and winter through spring 2006 using Pulse Amplitude Modulated (PAM) fluorometry including estimates of maximum quantum efficiency (F v/F m), maximum relative electron transport rate (rETRmax), photosynthetic efficiency (α), and the photoadaptive index (E k). The use of centrifuged brine samples allowed to document vertical gradients in ice algal acclimation with 5 cm vertical resolution for the first time. Bottom ice algae (0–5 cm from ice–water interface) expressed low F v/F m (0.331–0.426) and low α (0.098–0.130 (μmol photons m−2s−1)−1) in December. F v/F m and α increased in March and May (0.468–0.588 and 0.141–0.438 (μmol photons m−2s−1)−1, respectively) indicating increased photosynthetic activity. In addition, increases in rETRmax (3.3–16.4 a.u.) and E k (20–88 μmol photons m−2 s−1) from December to May illustrates a higher potential for primary productivity as communities become better acclimated to under-ice light conditions. In conclusion, photosynthetic performance by ice algae (as assessed by PAM fluorometry) was tightly linked to sea ice salinity, temperature, and inorganic nutrient concentrations (mainly nitrogen).  相似文献   

12.
The gene encoding an α-l-arabinofuranosidase that could biotransform ginsenoside Rc {3-O-[β-d-glucopyranosyl-(1–2)-β-d-glucopyranosyl]-20-O-[α-l-arabinofuranosyl-(1–6)-β-d-glucopyranosyl]-20(S)-protopanaxadiol} to ginsenoside Rd {3-O-[β-d-glucopyranosyl-(1–2)-β-d-glucopyranosyl]-20-O-β-d-glucopyranosyl-20(S)-protopanaxadiol} was cloned from a soil bacterium, Rhodanobacter ginsenosidimutans strain Gsoil 3054T, and the recombinant enzyme was characterized. The enzyme (AbfA) hydrolyzed the arabinofuranosyl moiety from ginsenoside Rc and was classified as a family 51 glycoside hydrolase based on amino acid sequence analysis. Recombinant AbfA expressed in Escherichia coli hydrolyzed non-reducing arabinofuranoside moieties with apparent K m values of 0.53 ± 0.07 and 0.30 ± 0.07 mM and V max values of 27.1 ± 1.7 and 49.6 ± 4.1 μmol min−1 mg−1 of protein for p-nitrophenyl-α-l-arabinofuranoside and ginsenoside Rc, respectively. The enzyme exhibited preferential substrate specificity of the exo-type mode of action towards polyarabinosides or oligoarabinosides. AbfA demonstrated substrate-specific activity for the bioconversion of ginsenosides, as it hydrolyzed only arabinofuranoside moieties from ginsenoside Rc and its derivatives, and not other sugar groups. These results are the first report of a glycoside hydrolase family 51 α-l-arabinofuranosidase that can transform ginsenoside Rc to Rd.  相似文献   

13.
Amyloid beta (Aβ) peptide accumulation has been demonstrated to play a central role in Alzheimer’s disease (AD). Substantial evidence indicates that protein nitrotyrosination contributes to Aβ-dependent neurotoxicity; however, the molecular mechanism is unknown. Recent research has shown that Aβ complexes with heme to form Aβ–heme, and increases the pseudo-peroxidase activity of heme. We found that Aβ–heme uses H2O2 and NO2 to cause nitration of enolase and synaptic proteins more effectively than heme. Thus, the increased peroxidase activity of Aβ–heme may be the molecular link between excess Aβ and the widespread protein nitration in AD. Interestingly, the site of enolase nitration that was catalyzed by Aβ–heme is different from that induced by heme. Moreover, the secondary structural perturbations of Aβ–heme-treated and heme-treated enolase are also different. These observations suggest that Aβ–heme targets specific amino acid sequences in enolase. Furthermore, our data show that Aβ–heme peroxidase activity is independent of the aggregation state of Aβ, suggesting an important role of soluble Aβ in addition to Aβ aggregates and oligomers in AD pathogenesis.  相似文献   

14.
The extracellular matrix component collagen type VI demonstrates potent growth-stimulatory effects and has been associated with aggressive tumour growth. Although, juvenile angiofibromas (JAs) often exhibit an aggressive growth pattern, the collagen type VI expression of this fibrovascular tumour has not been addressed so far. RT-PCR, Western blot analysis and immunohistochemistry were used in this study to analyse collagen type VI, type VI collagen receptor subunits (integrin α1, α2, α10, α11 and β1) and the type VI collagen receptor NG2 in JAs (N = 15) and nasal mucosa (NM, N = 8) samples. The mRNA expression of all three collagen type VI chains was found to be up-regulated significantly (P < 10−3–10−5, adjusted) in JAs compared to NM tissues. The Western blot analysis proved highly prominent collagen-type VI expression in JAs. The ApoTome technique revealed strong collagen-type VI signals in tumour endothelium. NG2 (P < 10−3, adjusted) and α11-integrin (P = 0.04, adjusted) showed a significantly higher mRNA expression levels in JAs than in NM samples. NG2, α1-, α2- and β1-intergin were located to tumour vessels, and additional stromal signals were observed for NG2 and α1-integrin in JAs. This study demonstrates a prominent collagen-type VI expression in JAs. The collagen-type VI may exert an important growth stimulus in this tumour.  相似文献   

15.
Although Saccharomyces cerevisiae can form petite mutants with deletions in mitochondrial DNA (mtDNA) (ρ) and can survive complete loss of the organellar genome (ρo), the genetic factor(s) that permit(s) survival of ρ and ρo mutants remain(s) unknown. In this report we show that a function associated with the F1-ATPase, which is distinct from its role in energy transduction, is required for the petite-positive phenotype of S. cerevisiae. Inactivation of either the α or β subunit, but not the γ, δ, or ɛ subunit of F1, renders cells petite-negative. The F1 complex, or a subcomplex composed of the α and β subunits only, is essential for survival of ρo cells and those impaired in electron transport. The activity of F1 that suppresses ρo lethality is independent of the membrane Fo complex, but is associated with an intrinsic ATPase activity. A further demonstration of the ability of F1 subunits to suppress ρo lethality has been achieved by simultaneous expression of S. cerevisiae F1α and γ subunit genes in Kluyveromyces lactis– which allows this petite-negative yeast to survive the loss of its mtDNA. Consequently, ATP1 and ATP2, in addition to the previously identified AAC2, YME1 and PEL1/PGS1 genes, are required for establishment of ρ or ρo mutations in S. cerevisiae. Received: 20 March 1999 / Accepted: 18 July 1999  相似文献   

16.
Aqueous methanol extracts of rattail fescue (Vulpia myuros) inhibited the growth of roots and shoots of cress (Lepidium sativum), lettuce (Lactuca sativa), alfalfa (Medicago sativa), timothy (Phleum pratense), Digitaria sanguinalis and Lolium multiflorum. Increasing the extract concentration increased the inhibition, suggesting that rattail fescue may have growth inhibitory substances and possess allelopathic potential. The aqueous methanol extract of rattail fescue was purified and two main inhibitory substances were isolated and identified by spectral data as (−)-3-hydroxy-β-ionone and (+)-3-oxo-α-ionol. Both substances inhibited root and shoot growth of cress at concentrations greater than 0.3 μM. The concentrations required for 50% growth inhibition on root and shoot growth of cress, lettuce, alfalfa, timothy, D. sanguinalis and L. multiflorum were 2.7–19.7 μM for (−)-3-hydroxy-β-ionone, and 2.1–34.5 μM for (+)-3-oxo-α-ionol. The concentration of (−)-3-hydroxy-β-ionone and (+)-3-oxo-α-ionol, respectively, in rattail fescue was 7.8 and 3.7 μg g−1 fresh weight. Considering the endogenous level and the inhibitory activity, (−)-3-hydroxy-β-ionone and (+)-3-oxo-α-ionol may work as allelopathic substances in rattail fescue through the growth inhibition of neighboring plant species.  相似文献   

17.
1H-3-Hydroxy-4-oxoquinaldine 2,4-dioxygenase (Hod), catalyzing cleavage of its heteroaromatic substrate to form carbon monoxide and N-acetylanthranilate, belongs to the α/β hydrolase fold family of enzymes. Analysis of protein variants suggested that Hod has adapted active-site residues of the α/β hydrolase fold for the dioxygenolytic reaction. H251 was recently shown to act as a general base to abstract a proton from the organic substrate. Residue S101, which corresponds to the nucleophile of the catalytic triad of α/β-hydrolases, presumably participates in binding the heteroaromatic substrate. H102 and residues located in the topological region of the triad’s acidic residue appear to influence O2 binding and reactivity. A tyrosine residue might be involved in the turnover of the ternary complex [HodH+–3,4-dioxyquinaldine dianion–O2]. Absence of viscosity effects and kinetic solvent isotope effects suggests that turnover of the ternary complex, rather than substrate binding, product release, or proton movements, involves the rate-determining step in the reaction catalyzed by Hod.  相似文献   

18.
A maltooligosaccharide-forming α-amylase was produced by a new soil isolate Bacillus subtilis KCC103. In contrast to other Bacillus species, the synthesis of α-amylase in KCC103 was not catabolite-repressed. The α-amylase was purified in one step using anion exchange chromatography after concentration of crude enzyme by acetone precipitation. The purified α-amylase had a molecular mass of 53 kDa. It was highly active over a broad pH range from 5 to 7 and stable in a wide pH range between 4 and 9. Though optimum temperature was 65–70 °C, it was rapidly deactivated at 70 °C with a half-life of 7 min and at 50 °C, the half-life was 94 min. The K m and V max for starch hydrolysis were 2.6 mg ml−1 and 909 U mg−1, respectively. Ca2+ did not enhance the activity and stability of the enzyme; however, EDTA (50 mM) abolished 50% of the activity. Hg2+, Ag2+, and p-hydroxymercurybenzoate severely inhibited the activity indicating the role of sulfydryl group in catalysis. The α-amylase displayed endolytic activity and formed maltooligosaccharides on hydrolysis of soluble starch at pH 4 and 7. Small maltooligosaccharides (D2–D4) were formed more predominantly than larger maltooligosaccharides (D5–D7). This maltooligosaccharide forming endo-α-amylase is useful in bread making as an antistaling agent and it can be produced economically using low-cost sugarcane bagasse.  相似文献   

19.
In order to further elucidate the, roles of DNA topoisomerase II (topo II) subtypes, α and β, as drug targets in chemotherapy, we have determined the enzyme levels in K562 cells selected for resistance to mitoxantrone (K562/Mxn), daunorubicin (K562/Dnr) and idarubicin (K562/Ida 20 and K562/Ida 60), as well as topo II-DNA complex formation, DNA damage and cytotoxicity, induced by topo II interactive agents, for example etoposide, teniposide, mitoxantrone and amsacrine. As compared to the parental cells, topo IIα/β protein levels in K562/Mxn, K562/Dnr, K562/Ida 20 and 60 lines, measured with Western blot, were 17/67%, 85/88, 24/31% and 10/7% respectively. DNA damage, determined by DNA unwinding technique, induced by teniposide and amsacrine correlated with both topo IIα/β protein levels (r 2=0.8/0.9,P=0.03/0.01 andr 2=0.8/0.9,P=0.04/0.01, respectively). Topo II-DNA complex formation induced by all studied drugs correlated with topo IIβ protein levels (r 2-range 0.8–0.9,P-range 0.01–0.04), while the correlation with topo IIα was weaker. Topo IIα/β protein levels tended to show an inverse correlation with the cytotoxicity of etoposide (r 2=−0.9/−0.7,P=0.01/0.06). The overall topo II-DNA complex formation correlated with drug-induced DNA damage (r 2=0.9,P=0.0001), whilst not with the cytotoxicity. Our findings indicate that both topo II isozymes are the targets of the antitumor agents studied, and of potential clinical relevance for prediction of treatment efficacy. They could play a role in tailored chemotherapy.  相似文献   

20.
The aggregation of the peptide amyloid-β (Aβ) to form amyloid plaques is a key event in Alzheimer’s disease. It has been shown that CuII can bind to soluble Aβ and influence its aggregation properties. Three histidines and the N-terminal amine have been proposed to be involved in its coordination. Here, for the first time, we show isothermal titration calorimetry (ITC) measurements of the CuII binding to Aβ16 and Aβ28, models of the soluble Aβ. Moreover, different spectroscopic methods were applied. The studies revealed new insights into these CuII–Aβ complexes: (1) ITC showed two CuII binding sites, with an apparent K d of 10−7 and 10−5 M, respectively; (2) the high-affinity site has a smaller enthalpic contribution but a larger entropic contribution than the low-affinity binding site; (3) azide did not bind to CuII in the higher-affinity binding site, suggesting the absence of a weak, labile ligand; (4) azide could bind to the CuII in the low-affinity binding site in Aβ28 but not in Aβ16; (5) 1H-NMR suggests that the carboxylate of aspartic acid in position 1 is involved in the ligation to CuII in the high-affinity binding site; (6) the pK a of 11.3 of tyrosine in position 10 was not influenced by the binding of 2 equivalents of CuII.Electronic Supplementary Material Supplementary material is available to authorized users in the online version of this article at .  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