首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
α-Conarachin and α-arachin are metabolized differently during germination. The amount of α-conarachin decreases without the appearance of antigenic intermediary compounds. α-Arachin retains its original antigenic structure and becomes more anodic during germination, but a number of intermediary proteins, differing widely in electrophoretic mobility, form up to 14 days germination. The electrophoretic shift of α-arachin, detectable after 1 day of germination, might be caused by progressive deamidation of glutamine and asparagine. Decreases in amide content of protein extracted after germination support the hypothesis and suggest that α-arachin provides the first source of ammonia to the new seedling.  相似文献   

2.
Subcellular fractions from the cotyledon obtained by differential and density gradient centrifugation, and extracts of total proteins from both cotyledon and axial tissues were analyzed by diethylaminoethyl cellulose chromatography, zone electrophoresis, ultracentrifugation, immunodiffusion, and immunoelectrophoresis. Fractionation and characterization of proteins in subcellular organelles of the peanut reaffirm that α-arachin is located in the protein bodies of the cells. Results obtained by diethylaminoethyl cellulose chromatography of subcellular fractions suggest that some of the conarachin proteins are cytoplasmic. α1-Conarachin is cytoplasmic, and α2-conarachin is particle-bound. α-Arachin and α2-conarachin predominate in the cotyledon. Quantitative differences for other proteins were also observed. Although qualitative similarities are apparent by immunoelectrophoresis, major differences were observed in the sedimentation patterns, zone electrophoreograms, and in the diethylaminoethyl cellulose chromatograms of total protein extracts from the cotyledon and the axis.  相似文献   

3.
Methionine residues in α/β-type small, acid-soluble spore proteins (SASP) of Bacillus species were readily oxidized to methionine sulfoxide in vitro by t-butyl hydroperoxide (tBHP) or hydrogen peroxide (H2O2). These oxidized α/β-type SASP no longer bound to DNA effectively, but DNA binding protected α/β-type SASP against methionine oxidation by peroxides in vitro. Incubation of an oxidized α/β-type SASP with peptidyl methionine sulfoxide reductase (MsrA), which can reduce methionine sulfoxide residues back to methionine, restored the α/β-type SASP’s ability to bind to DNA. Both tBHP and H2O2 caused some oxidation of the two methionine residues of an α/β-type SASP (SspC) in spores of Bacillus subtilis, although one methionine which is highly conserved in α/β-type SASP was only oxidized to a small degree. However, much more methionine sulfoxide was generated by peroxide treatment of spores carrying a mutant form of SspC which has a lower affinity for DNA. MsrA activity was present in wild-type B. subtilis spores. However, msrA mutant spores were no more sensitive to H2O2 than were wild-type spores. The major mechanism operating for dealing with oxidative damage to α/β-type SASP in spores is DNA binding, which protects the protein’s methionine residues from oxidation both in vitro and in vivo. This may be important in vivo since α/β-type SASP containing oxidized methionine residues no longer bind DNA well and α/β-type SASP-DNA binding is essential for long-term spore survival.  相似文献   

4.
Two ammonium-inducible, chloroplast-localized NADP-specific glutamate dehydrogenase isoenzymes were purified to homogeneity from Chlorella sorokiniana. These isoenzymes were homopolymers of either α- or β-subunits with molecular weights of 55,500 or 53,000, respectively. The α-isoenzyme was preferentially induced at low ammonium concentrations (2 millimolar or lower), whereas only the β-isoenzyme accumulated after cells were fully induced (120 minutes) at high ammonium concentrations (29 millimolar). Purification of isoenzymes was achieved by (NH4)2SO4 fractionation, gel-filtration, anion-exchange fast protein liquid chromatography, and affinity chromatography. The α- and β-isoenzymes were separated by their differential binding to Type 4 nicotinamide adenine dinucleotide phosphate-Sepharose. Both isoenzymes bound to an antibody affinity column to which purified antibody (prepared against β-isoenzyme) was covalently attached. Peptide mapping of the subunits showed them to have a high degree of sequence homology. Both subunits were synthesized in vitro from precursor protein(s) with a molecular weight of 58,500. Although the subunits have similar chemical, physical, and antigenic properties, their holoenzymes have strikingly different ammonium Km values. The ammonium Km of the β-isoenzyme remained constant at approximately 75 millimolar, whereas this Km of the α-isoenzyme ranged from 0.02 to 3.5 millimolar, depending upon nicotinamide adenine dinucleotide phosphate concentration.  相似文献   

5.
Lambda coli phage is not inactivated by chymotrypsin, trypsin, or ficin. T2 phage is slowly inactivated by high concentrations of (α-, β-, γ-, or Δ-chymotrypsin, but not by trypsin or ficin. P1 phage is slowly inactivated by α-, β-, or γ-chymotrypsin, or ficin, more rapidly by Δ-chymotrypsin, and much more rapidly by trypsin. Crystalline egg albumin, crystalline serum albumin, E. coli nucleoprotein, and yeast nucleoprotein are hydrolyzed slowly by α-chymotrypsin. Yeast nucleoprotein, like P1 phage, is hydrolyzed more rapidly by Δ-chymotrypsin than by α-chymotrypsin, but not by trypsin or ficin. Neither phages nor native proteins were attacked by papain, carboxypeptidase, deoxyribonuclease, or ribonuclease.  相似文献   

6.
The genome of Chlamydia trachomatis, one of the most prominent human pathogens, contains two structural genes coding for proteins, herein called Npt1Ct and Npt2Ct (nucleoside phosphate transporters 1 and 2 of C. trachomatis), exhibiting 68 and 61% similarity, respectively, to the ATP/ADP transporter from the intracellular bacterium Rickettsia prowazekii at the deduced amino acid level. Hydropathy analysis and sequence alignments suggested that both proteins have 12 transmembrane domains. The putative transporters were expressed as histidine-tagged proteins in Escherichia coli to study their biochemical properties. His10-Npt1Ct catalyzed ATP and ADP transport in an exchange mode. The apparent Km values were 48 (ATP) and 39 (ADP) μM. ATP and ADP transport was specific since AMP, GTP, CTP, UTP, dATP, dCTP, dGTP, and dTTP did not inhibit uptake. In contrast, His10-Npt2Ct transported all four ribonucleoside triphosphates with apparent Km values of 31 μM (GTP), 302 μM (UTP), 528 μM (CTP), and 1,158 μM (ATP). Ribonucleoside di- and monophosphates and deoxyribonucleotides were not substrates. The protonophore m-chlorocarbonylcyanide phenylhydrazone abolished uptake of all nucleoside triphosphates by Npt2Ct. This observation indicated that His10-Npt2Ct acts as a nucleosidetriphosphate/H+ symporter energized by the proton motive force across the Escherichia coli cytoplasmic membrane. We conclude that Npt1Ct provides chlamydiae with energy whereas Npt2Ct catalyzes the net uptake of ribonucleoside triphosphates required for anabolic reactions.  相似文献   

7.
Gibberellin A1 (GA1), 3-epi-GA1, GA4, GA9, 11α-hydroxyGA12, 12α-hydroxyGA12, GA15, GA17, GA19, GA20, GA25, GA37, GA40, GA58, GA69, GA70, and GA71 have been identified from Kovats retention indices and full scan mass spectra by capillary GC-MS analyses of purified extracts from sporophytes of the tree fern, Cibotium glaucum. Abscisic acid, dihydrophaseic acid, an epimer of 4′-dihydrophaseic acid, and the epimeric ent-6α, 7α, 16α, 17-(OH)4 and ent-6α, 7α, 16β, 17-(OH)4 derivatives of ent16, 17-dihydrokaurenoic acid, in addition to the epimeric 16α, 17- and 16β, 17-dihydroxy-16, 17-dihydro derivatives of GA12, were also identified in extracts of C. glaucum. An oxodihydrophaseic acid and a hydroxydihydrophaseic acid were also detected. In extracts of sporophytes of Dicksonia antarctica, GA4, GA9, 12α- and 12β-hydroxyGA12, GA15, GA25, and GA37 were identified by the same criteria, as well as abscisic acid, phaseic acid, 8′-hydroxymethylabscisic acid and dihydrophaseic acid. This is the first time that GA40 has been identified in a higher plant; it is also the first report of the natural occurrence of the two gibberellins, 11α- and 12β-hydroxyGA12. The total gibberellin (GA) content in C. glaucum (tall) was at least one order of magnitude greater than that of D. antarctica (dwarf) based on total ion current response in GC-MS and bioassay data. Abscisic acid was a major component of D. antarctica and the oxodihydrophaseic acid was a major component of C. glaucum.  相似文献   

8.
The phase transitions of dipalmitoylphosphatidylethanolamine (DPPE) in excess water have been examined by low-angle time-resolved x-ray diffraction and calorimetry at low scan rates. The lamellar subgel/lamellar liquid-crystalline (Lc → Lα), lamellar gel/lamellar liquid-crystalline (Lβ → Lα), and lamellar liquid-crystalline/lamellar gel (Lα → Lβ) phase transitions proceed via coexistence of the initial and final phases with no detectable intermediates at scan rates 0.1 and 0.5°C/min. At constant temperature within the region of the Lβ → Lα transition the ratio of the two coexisting phases was found to be stable for over 30 min. The state of stable phase coexistence was preceded by a 150-s relaxation taking place at constant temperature after termination of the heating scan in the transition region. While no intermediate structures were present in the coexistence region, a well reproducible multipeak pattern, with at least four prominent heat capacity peaks separated in temperature by 0.4-0.5°C, has been observed in the cooling transition (Lα → Lβ) by calorimetry. The multipeak pattern became distinct with an increase of incubation time in the liquid-crystalline phase. It was also clearly resolved in the x-ray diffraction intensity versus temperature plots recorded at slow cooling rates. These data suggest that the equilibrium state of the Lα phase of hydrated DPPE is represented by a mixture of domains that differ in thermal behavior, but cannot be distinguished structurally by x-ray scattering.  相似文献   

9.
A complex of the three (αεθ) core subunits and the β2 sliding clamp is responsible for DNA synthesis by Pol III, the Escherichia coli chromosomal DNA replicase. The 1.7 Å crystal structure of a complex between the PHP domain of α (polymerase) and the C-terminal segment of ε (proofreading exonuclease) subunits shows that ε is attached to α at a site far from the polymerase active site. Both α and ε contain clamp-binding motifs (CBMs) that interact simultaneously with β2 in the polymerization mode of DNA replication by Pol III. Strengthening of both CBMs enables isolation of stable αεθ:β2 complexes. Nuclear magnetic resonance experiments with reconstituted αεθ:β2 demonstrate retention of high mobility of a segment of 22 residues in the linker that connects the exonuclease domain of ε with its α-binding segment. In spite of this, small-angle X-ray scattering data show that the isolated complex with strengthened CBMs has a compact, but still flexible, structure. Photo-crosslinking with p-benzoyl-L-phenylalanine incorporated at different sites in the α-PHP domain confirm the conformational variability of the tether. Structural models of the αεθ:β2 replicase complex with primer-template DNA combine all available structural data.  相似文献   

10.
W. L. Hardy 《Biophysical journal》1973,13(10):1054-1070
Conduction speed (θ) in single myelinated Rana pipiens sciatic nerve fibers has been precisely measured using intracellular recording and on-line digital computer techniques. The dependence of relative speed on external Na concentration at 15°C has been found to be ln(θ12) = 0.524 (±0.018) ln ([Na+]1/[Na+]2) + 0.003. Thus θ has very close to a square root dependence on [Na+]0 for these fibers. This experimental finding is not in complete agreement with a theoretical prediction based on a solution of the Hodgkin-Huxley (H.H.) equations. The effect of small temperature variations around 15°C on θ has also been measured for Rana fibers in Ringer's solution. θ has close to an exponential dependence on T and a Q10 of 2.95 has been estimated.  相似文献   

11.
1. The subunits α and β of Halobacterium cutirubrum DNA-dependent RNA polymerase have been purified to electrophoretic homogeneity. Both have mol.wt. 18000 and they are required in equimolar amounts for optimum activity. 2. The instability of the complete enzyme, αβ, in the absence of salt is due to the rapid inactivation of the β subunit in these conditions. 3. Nearest-neighbour analysis of the product formed on poly[d(A-T)] as template shows that the enzyme copies the latter accurately. 4. The enzyme initiates new chains with purine nucleoside triphosphates exclusively. 5. The product obtained in the standard assay conditions contains some high mol.wt. (>16S) material, but consists primarily of short chains, of average length 70–80 nucleotide units. 6. The template specificity of the complete enzyme has been studied at high and low ionic strength. Its extreme dependence on salt concentration is unrelated to the gross overall base composition of the DNA used. 7. T7 DNA is transcribed asymmetrically and the enzyme selectively copies the T7 `early' genes. 8. Preliminary amino acid analyses of α and β subunits show that their overall content of acidic, basic and neutral amino acids does not differ appreciably from that of Escherichia coli RNA polymerase.  相似文献   

12.
Yang Y  Yin C  Li W  Xu X 《Journal of bacteriology》2008,190(5):1554-1560
Unlike Escherichia coli, the cyanobacterium Synechocystis sp. strain PCC 6803 is insensitive to chill (5°C) in the dark but rapidly losses viability when exposed to chill in the light (100 μmol photons m−2 s−1). Preconditioning at a low temperature (15°C) greatly enhances the chill-light tolerance of Synechocystis sp. strain PCC 6803. This phenomenon is called acquired chill-light tolerance (ACLT). Preconditioned wild-type cells maintained a substantially higher level of α-tocopherol after exposure to chill-light stress. Mutants unable to synthesize α-tocopherol, such as slr1736, slr1737, slr0089, and slr0090 mutants, almost completely lost ACLT. When exposed to chill without light, these mutants showed no or a slight difference from the wild type. When complemented, the slr0089 mutant regained its ACLT. Copper-regulated expression of slr0090 from PpetE controlled the level of α-tocopherol and ACLT. We conclude that α-tocopherol is essential for ACLT of Synechocystis sp. strain PCC 6803. The role of α-tocopherol in ACLT may be based largely on a nonantioxidant activity that is not possessed by other tocopherols or pathway intermediates.  相似文献   

13.
Antibodies were raised against carrot (Daucus carota) cell wall β-fructosidase that was either in a native configuration (this serum is called anti-βF1) or chemically deglycosylated (anti-βF2). The two antisera had completely different specificities when tested by immunoblotting. The anti-βF1 serum reacted with β-fructosidase and many other carrot cell wall proteins as well as with many proteins in extracts of bean (Phaseolus vulgaris) cotyledons and tobacco (Nicotiana tabacum) seeds. It did not react with chemically deglycosylated β-fructosidase. The anti-βF1 serum also reacted with the bean vacuolar protein, phytohemagglutinin, but not with deglycosylated phytohemagglutinin. The anti-βF2 serum reacted with both normal and deglycosylated β-fructosidase but not with other proteins. These results indicate that the βF2 antibodies recognize the β-fructosidase polypeptide, while the βF1 antibodies recognize glycan sidechains common to many glycoproteins. We used immunoadsorption on glycoprotein-Sepharose columns and hapten inhibition of immunoblot reactions to characterize the nature of the antigenic site. Antibody binding activity was found to be associated with Man3(Xyl)(GIcNAc)2Fuc, Man3(Xyl)(GIcNAc)2, and Man(Xyl) (GIcNAc)2 glycans, but not with Man3(GIcNAc)2. Treatment of phytohemagglutinin, a glycoprotein with a Man3(Xyl)(GIcNAc)2Fuc glycan, with Charonia lampas β-xylosidase (after treatment with jack-bean α-mannosidase) greatly diminished the binding between the antibodies and phytohemagglutinin. We conclude, therefore, that the antibodies bind primarily to the xyloseβ, 1→ 2mannose structure commonly found in the complex glycans of plant glycoproteins.  相似文献   

14.
1. A fragment of human serum albumin called `inhibitor' has been degraded by trypsin, and one of the degradation products, designated fragment F1, has been isolated. Fragment F1 has a molecular weight of 6600. It contains neither tyrosine nor tryptophan. It is not precipitated with rabbit anti-sera to human serum albumin. 2. Fragment F1 was coupled to p-aminobenzylcellulose to form an insoluble conjugate. Rabbit anti-(human serum albumin) antibodies reacting with fragment F1 were specifically adsorbed on this conjugate and were desorbed by glycine–hydrochloric acid buffer. The isolated antibodies are composed of γ-globulin and β2-macroglobulin. 3. Human serum albumin and fragment F1 formed with 7s anti-(fragment F1) antibodies soluble complexes that were studied by passive haemagglutination, ultracentrifugation and electrophoresis. Fragment F1 was shown to contain only one of the antigenic sites of albumin molecule. The 7s anti-(fragment F1) antibodies were shown to be bivalent and monospecific.  相似文献   

15.
There are few entries of carbon-carbon bond hydrolases (EC 3.7.1.-) in the ExPASy database. In microbes, these enzymes play an essential role in the metabolism of alicyclic or aromatic compounds as part of the global carbon cycle. CpdC is a ω-pentadecalactone hydrolase derived from the degradation pathway of cyclopentadecanol or cyclopentadecanone by Pseudomonas sp. strain HI-70. CpdC was purified to homogeneity and characterized. It is active as a dimer of 56,000 Da with a subunit molecular mass of 33,349. Although CpdC has the highest activity and reaction rate (kcat) toward ω-pentadecalactone, its catalytic efficiency favors lauryl lactone as a substrate. The melting temperature (Tm) of CpdC was estimated to be 50.9 ± 0.1°C. The half-life of CpdC at 35°C is several days. By virtue of its high level of expression in Escherichia coli, the intact CpdC-encoding gene and progressive 3′-end deletions were employed in the construction of a series of fusion plasmid system. Although we found them in inclusion bodies, proof-of-concept of overproduction of three microbial cutinases of which the genes were otherwise expressed poorly or not at all in E. coli was demonstrated. On the other hand, two antigenic proteins, azurin and MPT63, were readily produced in soluble form.  相似文献   

16.
n-Alkyl (C6-C12) β-d-monoglucopyranosides have been found to be highly potent activators of mung bean β-glucan synthase in vitro, increasing the Vmax of the enzyme as much as 60-fold and with Ka values as low as 10 micromolar. Activation is highly specific for the β-linked terminal glucose residue; other alkyl glycosides such as, octyl-α-glucoside, dodecyl β-maltoside, 6-lauryl sucrose, 6-lauryl glucose, which lack this structure, are ineffective as activators. Based on the similarities in their structure and effects on β-glucan synthesis under a variety of conditions, it is proposed that the alkyl β-glucosides are structural analogs of the native glucolipid activator of β-glucan synthase isolated from mung bean extracts.  相似文献   

17.
The elastic behavior of the cell wall as a function of the temperature has been studied with particular attention being given to the swelling of egg cells of Strongylocentrotus purpuratus and Crassostrea virginica in different sea water concentrations at different temperatures. It was found that the modulus of elasticity is a nonlinear function of temperature. At about 12-13°C the modulus of elasticity (E) is constant, independent of the stress (σ) and strain (εν) which exist at the cell wall; the membranous material follows Hooke's law, and E ≈ 3 × 107 dyn/cm2 for S. purpuratus and C. virginica. When the temperature is higher or lower than 12-13°C, the modulus of elasticity increases, and the membranous material does not follow Hooke's law, but is almost directly proportional to the stresses existing at the cell wall. On increasing the stress, the function Eσ = E(σ) approaches saturation. The corresponding stress-strain diagrams, σ = σ(εν), and the graphs, Eσ = E(σ) and Eσ = E(t) are given. The cyto-elastic phenomena at the membrane are discussed.  相似文献   

18.
Single-molecule pulling experiments on unstructured proteins linked to neurodegenerative diseases have measured rupture forces comparable to those for stable folded proteins. To investigate the structural mechanisms of this unexpected force resistance, we perform pulling simulations of the amyloid β-peptide (Aβ) and α-synuclein (αS), starting from simulated conformational ensembles for the free monomers. For both proteins, the simulations yield a set of rupture events that agree well with the experimental data. By analyzing the conformations occurring shortly before rupture in each event, we find that the mechanically resistant structures share a common architecture, with similarities to the folds adopted by Aβ and αS in amyloid fibrils. The disease-linked Arctic mutation of Aβ is found to increase the occurrence of highly force-resistant structures. Our study suggests that the high rupture forces observed in Aβ and αS pulling experiments are caused by structures that might have a key role in amyloid formation.  相似文献   

19.
Self-assembly of the intrinsically unstructured proteins, amyloid beta (Aβ) and alpha synclein (αSyn), are associated with Alzheimer’s Disease, and Parkinson’s and Lewy Body Diseases, respectively. Importantly, pathological overlaps between these neurodegenerative diseases, and the possibilities of interactions between Aβ and αSyn in biological milieu emerge from several recent clinical reports and in vitro studies. Nevertheless, there are very few molecular level studies that have probed the nature of spontaneous interactions between these two sequentially dissimilar proteins and key characteristics of the resulting cross complexes. In this study, we have used atomistic molecular dynamics simulations to probe the possibility of cross dimerization between αSyn1–95 and Aβ 1–42, and thereby gain insights into their plausible early assembly pathways in aqueous environment. Our analyses indicate a strong probability of association between the two sequences, with inter-protein attractive electrostatic interactions playing dominant roles. Principal component analysis revealed significant heterogeneity in the strength and nature of the associations in the key interaction modes. In most, the interactions of repeating Lys residues, mainly in the imperfect repeats ‘KTKEGV’ present in αSyn1–95 were found to be essential for cross interactions and formation of inter-protein salt bridges. Additionally, a hydrophobicity driven interaction mode devoid of salt bridges, where the non-amyloid component (NAC) region of αSyn1–95 came in contact with the hydrophobic core of Aβ 1–42 was observed. The existence of such hetero complexes, and therefore hetero assembly pathways may lead to polymorphic aggregates with variations in pathological attributes. Our results provide a perspective on development of therapeutic strategies for preventing pathogenic interactions between these proteins.  相似文献   

20.
Two α4β2 nicotinic acetylcholine receptor (α4β2-nAChR) isoforms exist with (α4)2(β2)3 and (α4)3(β2)2 subunit stoichiometries and high versus low agonist sensitivities (HS and LS), respectively. Both isoforms contain a pair of α4(+)/(−)β2 agonist-binding sites. The LS isoform also contains a unique α4(+)/(−)α4 site with lower agonist affinity than the α4(+)/(−)β2 sites. However, the relative roles of the conserved α4(+)/(−)β2 agonist-binding sites in and between the isoforms have not been studied. We used a fully linked subunit concatemeric nAChR approach to express pure populations of HS or LS isoform α4β2*-nAChR. This approach also allowed us to mutate individual subunit interfaces, or combinations thereof, on each isoform background. We used this approach to systematically mutate a triplet of β2 subunit (−)-face E-loop residues to their non-conserved α4 subunit counterparts or vice versa (β2HQT and α4VFL, respectively). Mutant-nAChR constructs (and unmodified controls) were expressed in Xenopus oocytes. Acetylcholine concentration-response curves and maximum function were measured using two-electrode voltage clamp electrophysiology. Surface expression was measured with 125I-mAb 295 binding and was used to define function/nAChR. If the α4(+)/(−)β2 sites contribute equally to function, making identical β2HQT substitutions at either site should produce similar functional outcomes. Instead, highly differential outcomes within the HS isoform, and between the two isoforms, were observed. In contrast, α4VFL mutation effects were very similar in all positions of both isoforms. Our results indicate that the identity of subunits neighboring the otherwise equivalent α4(+)/(−)β2 agonist sites modifies their contributions to nAChR activation and that E-loop residues are an important contributor to this neighbor effect.  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