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 222 毫秒
1.
Pseudomonas resinovorans phaC1 Pre and phaC2 Pre genes coding for poly(hydroxyalkanoate) (PHA) synthases were cloned by PCR and expressed in E. coli LS1298 (fadB). Repeat-unit composition analysis showed that -hydroxydecanoate (67–75 mol%) and -hydroxyoctanoate (25–33 mol%) are the major monomers of the PHA produced in cells grown on decanoate. Sequence analysis showed that the gene products of phaC1 Pre and phaC2 Pre had 61% identical (75% positive) amino-acid sequence matches, and both sequences contained a conserved /-hydrolase fold in the carboxy-terminal portion of the proteins. Switching the /-hydrolase folds of phaC1 Pre and phaC2 Pre yielded chimeric pha7 and pha8 genes that afforded PHA synthesis in E. coli LS1298. The repeat-unit compositions of PHA in cells containing pha7 and pha8 were similar to those found in transformants containing the parental genes. Deletion mutants of phaC1 Pre and phaC2 Pre that resulted in potential translational fusions also supported PHA synthesis with similar repeat-unit compositions. Chimeric genes obtained from the switching of fragments containing the /-hydrolase folds of phaC1 Pre and Ralstonia eutropha phbC did not direct the synthesis of PHA in transformed cells.  相似文献   

2.
Recombinant strains of Wautersia eutropha expressing an artificial polyhydroxyalkanoate (PHA) biosynthesis operon under the control of different native promoters linked to polyhydroxybutyrate (PHB) (Pphb), acetoin (PacoE, PacoD, and PacoX) or pyruvate (PpdhE) metabolism were constructed and tested. The promoters were representative either of the enterobacterial σ70 (Pphb, PacoE, and PpdhE)- or σ54 (PacoD and PacoX)-dependent promoters. To obtain polymers consisting of C4–C12 monomer units, an artificial operon consisting of the PHA synthase gene from Pseudomonas sp. 61-3 (phaC1 Ps) tandemly linked to the W. eutropha genes encoding β-ketothiolase (phbA We) and nicotinamide adenine dinucleotide phosphate dependent acetoacetyl-coenzyme A (CoA) reductase (phbB We) was constructed. All recombinant strains produced PHA, indicating that the PHA biosynthesis genes were expressed under the control of the different promoters. Cell growth and PHA synthesis on MS medium complemented with gluconate or octanoate, and different concentrations of acetoin (0, 0.15, and 0.3%) clearly differed among the recombinant strains. While the PacoD and PacoX promoters mediated only low PHA yields (<1%) in the presence of the inducer acetoin, the remaining promoters—independent of the addition of acetoin—resulted in the production of PHA polymers with high 3HB fractions (90–100 mol%) and with high 3HO contents (70–86 mol%) from gluconate and octanoate, respectively. Interestingly, on octanoate-MS medium with 0.15% acetoin, the PacoE promoter mediated the synthesis of PHA with a relatively high 3HB fraction (48 mol%). While PHAs with high 3HB contents were obtained, the overall PHA product yields were low (<10%); thus, their potential application for further commercial exploitation appears limited.  相似文献   

3.
Medium optimization for production of poly-β-hydroxyalkanoate (PHA) from Rhodobacter sphaeroides U7 cultivated in glutamate–acetate (GA) medium supplemented with 40 mM valeric acid as co-substrate under aerobic-dark condition was investigated. Studies on effect of nitrogen source and cultivation temperature by conventional and statistical methods illustrated that (NH4)2SO4 (0.2 g/l) had no effect and the optimal temperature was at 30°C. The optimum environmental conditions were found to be anaerobic-light (3000 lux) cultivation with aeration rate of 1.0 vvm and agitation speed of 200 rpm for PHA production (2.5 g/l) with the highest PHA content (65.15%) at 0.5 vvm, and 200 rpm. Under this optimized medium and condition, PHA production from R. sphaeroides U7 increased 3.86-folds (from 0.69 to 2.66 g/l) (PHA content increased 1.5-folds). The biopolymer was purified and characterized by using 13C NMR, FTIR, DSC, X-ray diffraction and intrinsic viscosity techniques to be a copolymer poly(β-hydroxybutyrate-co-β-hydroxyvalerate) (PHBV) consisting of 84.8 mol% β-hydroxybutyric acid (HB) and 15.2 mol% β-hydroxyvaleric acid (HV).  相似文献   

4.
 The dynamic quenching of the luminescence of racemic Eu(III)(pyridine-2,6-dicarboxylate=dpa)3 3– by the title proteins is investigated and the enantioselectivity of the proteins in the quenching of the Δ and Λ enantiomers of Eu(dpa)3 3– is determined. The two diastereomeric quenching rate constants pertaining to azurin (k q Δ=3.3×106, k q Λ=2.7×106 M–1 s–1, pH 7.2, ionic strength I=22 mM) are lower than for its Met→44Lys mutant (k q Δ=1.9×107, k q Λ=1.4×107 M–1 s–1, same pH and I), indicating that energy transfer occurs from Eu(dpa)3 3– to the Cu(II) centre when the luminophore is bound to the hydrophobic patch of the protein near residue 44. The enantioselectivity remains unaltered by the mutation: k q Δ/k q Λ=1.27±0.04, so Lys44 is probably not in direct contact with the Eu chelate. The I and pH dependence of k q indicate that the lysine residue interacts electrostatically with Eu(dpa)3 3–. For plastocyanin the quenching rates are of the order of 106 M–1 s–1; for amicyanin they are two orders of magnitude larger (k q Δ=12×107, k q Λ=11×107 M–1 s–1, pH 7.2, I=22 mM). The variation of k q is attributed to differences in the charge distribution on the proteins, which influences the binding of the luminophore to the protein surface. For amicyanin the anion binding site near Lys59 and Lys60 may be involved in the energy transfer. Received: 16 June 1998 / Accepted: 18 September 1998  相似文献   

5.
By the use of the Immobiline technique at pH ranges 7.0–7.6 and 6.9–7.9, 16 different hemoglobin (Hb) phenotypes were observed in 61 English Saanen goats. They are explained in this breed by a genetic theory of five β-globin genes (A 4,A 6,A 8,E, andD) and two closely linked α-globin loci (′α and ″α) of which the ″α has a variant allele, provisionally called ″α X . Family data together with observed and expected Hb frequencies were in agreement with the genetic theory. Among six Barbary sheep there were three Hb phenotypes explained by the occurrence of the β-chain allelesB andC na.  相似文献   

6.
The effect of mutations indnaK anddnaJ genes on the expression of two operons that are part of cysteine regulon was determined usingEscherichia coli strains harboringcysPTWA::lacZ andcysJIH::lacZ fusions. NulldnaJ, anddnaKdnaJ mutants were impaired in β-galactosidase expression from both fusions. Effecient complementation of this defect by wild-type alleles present on a low-copy number plasmid was achieved. The presence of the pMH224 plasmid coding for CysB* protein defective in DNA binding lowered β-galactosidase expression fromcysPTWA::lacZ fusion strain harboring wild-typednaKdnaJ alleles but did not diminish enzyme expression in ΔdnaJ and ΔdnaKdnaJ strains.  相似文献   

7.
Summary. For-Met-βAlaψ[CSNH]-Phe-OMe (3), For-Met-βAlaψ[CH2NH]-Phe-OMe (5), For-Met-NH-pC6H4-SO2-Phe-OMe (8a), For-Met-NH-mC6H4-SO2-Phe-OMe (8b) and the corresponding N-Boc precursors (2, 4, 7a, b) have been synthesized and their activity towards human neutrophils has been evaluated in comparison with that shown by the reference tripeptide For-Met-Leu-Phe-OMe (fMLF-OMe). Chemotaxis, lysozyme release and superoxide anion production have been measured. 1H NMR titration experiments and IR spectra have been discussed in order to ascertain the preferred solution conformation adopted by the tripeptide 3 with particular reference to the presence of a folded conformation centred at the centrally positioned thionated β-residue.  相似文献   

8.
Diphtheria toxin (DT) forms cation selective channels at low pH in cell membranes and planar bilayers. The channels formed by wild-type full length toxin (DT-AB), wild-type fragment B (DT-B) and mutants of DT-B were studied in the plasma membrane of Vero cells using the patch-clamp technique. The mutations concerned certain negatively charged amino acids within the channel-forming transmembrane domain (T-domain). These residues might interact electrostatically with cations flowing through the channel, and were therefore exchanged for uncharged amino acids or lysine. The increase in whole-cell conductance induced by toxin, Δg m , was initially determined. DT-AB induced a ∼10-fold lower Δg m than DT-B. The mutations DT-B E327Q, DT-B D352N and DT-B E362K did not affect Δg m , whereas DT-B D295K, DT-B D352K and DT-B D318K drastically reduced Δg m . Single channel analysis of DT-B, DT-AB, DT-B D295K, DT-B D318K and DT-B E362K was then performed in outside-out patches. No differences were found for the single-channel conductances, but the mutants varied in their gating characteristics. DT-B D295K exhibited only a very transient channel activity. DT-AB as well as DT-B D318K displayed significantly lower open probability and mean dwell times than DT-B. Hence, the lower channel forming efficiency of DT-AB and DT-B D318K as compared to DT-B is reflected on the molecular level by their tendency to spend more time in the closed position and the fast flickering mode. Altogether, the present work shows that replacements of single amino acids distributed throughout a large part of the transmembrane domain (T-domain) strongly affect the overall channel activity expressed as Δg m and the gating kinetics of single channels. This indicates clearly that the channel activity observed in DT-exposed Vero cells at low pH is inherent to DT itself and not due to DT-activation of an endogenous channel. Received: 20 June 1996/Revised: 8 November 1996  相似文献   

9.
 Direct cyclic voltammetry and 1H NMR spectroscopy have been combined to investigate the electrochemical and spectroscopic properties of cytochrome c 553 isolated from the alkaliphilic soil bacterium Bacillus pasteurii. A quasi-reversible diffusion-controlled redox process is exhibited by cytochrome c 553 at a pyrolitic graphite edge microelectrode. The temperature dependence of the reduction potential, measured using a non-isothermal electrochemical cell, revealed a discontinuity at 308 K. The thermodynamic parameters determined in the low-temperature range (275–308 K;ΔS°′=–162.7±1.2 J mol–1 K–1, ΔH°′=–53.0±0.5 kJ mol–1, ΔG°′=–4.5±0.1 kJ mol–1, E°′=+47.0±0.6 mV) indicate the presence of large enthalpic and entropic effects, leading, respectively, to stabilization and destabilization of the reduced form of cytochrome c 553. Both effects are more accentuated in the high-temperature range (308–323 K;ΔS°′=–294.1±8.4 J mol–1 K–1, ΔH°′=–93.4±3.1 kJ mol–1, ΔG°′=–5.8±0.6 kJ mol–1, E°′=+60.3±5.8 mV), with the net result being a slight increase of the standard reduction potential. These thermodynamic parameters are interpreted using the compensation theory of hydration of biopolymers as indicating the extrusion, upon reduction, of water molecules from the hydration sphere of the cytochrome. The low-T and high-T conformers differ by the number of water molecules in the solvation sphere: in the high-T conformer, the number of water molecules extruded upon reduction increases, as compared to the low-T conformer. The ionic strength dependence of the reduction potential at 298 K, treated within the frame of extended Debye-Hückel theory, yields values of E °′ (I=0) =–25.4±1.4 mV, z red=–11.3, and z ox=–10.3. The pH dependence of the reduction potential at 298 K shows a plateau in the pH range 7–10 and an increase at more acidic pH, allowing the calculation of pK O=5.5 and pK R=5.7, together with the estimate of the reduction potentials of completely protonated (+71 mV) and deprotonated (+58 mV) forms of cytochrome c 553. 1H NMR spectra of the oxidized paramagnetic cytochrome c 553 indicate the presence of a His-Met axial coordination of the low-spin (S=1/2) heme iron, which is maintained in the temperature interval 288–340 K at pH 7 and in the pH range 4.8–10.0 at 298 K. The temperature dependence of the hyperfine-shifted signals shows both Curie-type and anti-Curie-type behavior, with marked deviations from linearity, interpreted as indicating the presence of a fast equilibrium between the low-T and high-T conformers, having slightly different heme electronic structures resulting from the T-induced conformational change. Increasing the NaCl concentration in the range 0–0.2 M causes a slight change of the 1H NMR chemical shifts of the hyperfine-shifted signals, with no influence on their linewidth. The calculated lower limit value of the apparent affinity constant for specific ion binding is estimated as 5.2±1.1 M–1. The pH dependence of the isotropically shifted 1H NMR signals of the oxidized cytochrome displays at least one ionization step with pK O=5.7. The thermodynamic and spectroscopic data indicate a large solvent-derived entropic effect as the main cause for the observed low reduction potential of B. pasteurii cytochrome c 553. Received: 9 January 1998 / Accepted: 8 April 1998  相似文献   

10.
Using constant heat sap flow sensors, xylem water fluxes in ten tree species and two liana species were monitored for 5–10 days during the beginning of the wet season in May, 1993. For a subset of the trees, a branch was also monitored at the top of the crown for 5 days. Xylem flux (J S) was related diurnally in all plants to vapor pressure deficit (D) measured within the upper-third of the canopy, and to incoming shortwave radiation R S above the canopy. Cross-correlation analysis was used to estimate time lags between diurnal patterns of J S and D or R S, and between J S in stems and branches. The maximum correlation coefficient from cross-correlation of J S with R S (range=0.57–0.92) was often higher than the maximum of J S with D (range=0.43–0.89), indicating that diurnal J S was more dependent on R S than D. Time lags (lag corresponding to maximum correlation) of J S at stem-base with D was shorter (0–45 min) than with radiation (5–115 min), highly variable within a species, and uncorrelated to the height or exposure of tree crowns or liana in the canopy. On a stand level, not accounting for the diel lag between stem sap flux and canopy flux resulted in errors in estimated canopy transpiration of up to 30%. Received: 19 October 1998 / Accepted: 8 June 1999  相似文献   

11.
Comparative studies of the transglycosylation and hydrolytic activities have been performed on the Rhodothermus marinus β-1,3-glucanase (laminarinase) and its M133A, M133C, and M133W mutants. The M133C mutant demonstrated near 20% greater rate of transglycosylation activity in comparison with the M133A and M133W mutants that was measured by NMR quantitation of nascent β(1-4) and β(1-6) linkages. To obtain kinetic probes for the wild-type enzyme and Met-133 mutants, p-nitrophenyl β-laminarin oligosaccharides of degree of polymerisation 2–8 were synthesized enzymatically. Catalytic efficiency values, k cat/K m, of the laminarinase catalysed hydrolysis of these oligosaccharides suggested possibility of four negative and at least three positive binding subsites in the active site. Comparison of action patterns of the wild-type and M133C mutant in the hydrolysis of the p-nitrophenyl-β-D-oligosac- charides indicated that the increased transglycosylation activity of the M133C mutant did not result from altered subsite affinities. The stereospecificity of the transglycosylation reaction also was unchanged in all mutants; the major transglycosylation products in hydrolysis of p-nitrophenyl laminaribioside were β-glucopyranosyl-β-1,3-D-glucopy- ranosyl-β-1,3-D-glucopyranose and β-glucopyranosyl-β-1, 3-D-glucopyranosyl-β-1,3-D-glucpyranosyl-β-1,3-D- glucopyranoxside. In a memoriam of Dr. Kirill N. Neustroev. All we, his friends and colleagues, mourn for his sudden death. He was a bright and talented scientist, brilliant manager and good friend.  相似文献   

12.
Pseudomonas sp. 61-3 accumulated a blend of poly(3-hydroxybutyrate) [P(3HB)] homopolymer and a random copolymer consisting of 3-hydroxyalkanoate (3HA) units of 4–12 carbon atoms. The genes encoding β-ketothiolase (PhbARe) and NADPH-dependent acetoacetyl-CoA reductase (PhbBRe) from Ralstoniaeutropha were expressed under the control of promoters for Pseudomonas sp. 61-3 pha locus or R. eutropha phb operon together with phaC1 Ps gene (PHA synthase 1 gene) from Pseudomonas sp. 61-3 in PHA-negative mutants P. putida GPp104 and R. eutropha PHB4 to produce copolyesters [P(3HB-co-3HA)] consisting of 3HB and medium-chain-length 3HA units of 6–12 carbon atoms. The introduction of the three genes into GPp104 strain conferred the ability to synthesize P(3HB-co-3HA) with relatively high 3HB compositions (up to 49 mol%) from gluconate and alkanoates, although 3HB units were not incorporated at all or at a very low fraction (3 mol%) into copolyesters by the strain carrying phaC1 Ps gene only. In addition, recombinant strains of R. eutropha PHB4 produced P(3HB-co-3HA) with higher 3HB fractions from alkanoates and plant oils than those from recombinant GPp104 strains. One of the recombinant strains, R. eutropha PHB4/pJKSc46-pha, in which all the genes introduced were expressed under the control of the native promoter for Pseudomonas sp. 61-3 pha locus, accumulated P(3HB-co-3HA) copolyester with a very high 3HB fraction (85 mol%) from palm oil. The nuclear magnetic resonance analyses showed that the copolyesters obtained here were random copolymers of 3HB and 3HA units. Received: 12 July 1999 / Received revision: 1 October 1999 / Accepted: 2 October 1999  相似文献   

13.
Summary α,β-Dehydroamino acids are useful peptide modifiers. However, their stereoelectronic properties still remain insufficiently recognized. Based on FTIR experiments in the range ofv s(N-H), AI, AII andv s(Cα=Cβ) and ab initio calculations with B3LYP/6–31G*, we studied the solution conformational preferences and the amide electron density perturbation of Ac-ΔXaa-NHMe, where ΔXaa=ΔAla, (E)-ΔAbu, (Z)-ΔAbu, (Z)-ΔLeu, (Z)-ΔPhe and ΔVal. Each of these dehydroamides adopts a C5 structure, which in Ac-ΔAla-NHMe is fully extended and accompanied by the strong C5 hydrogen bond. Interaction with bond Cα=Cβ lessens the amidic resonance within the flanking amide groups. TheN-terminal C=O bond is noticeably shorter, both amide bonds are longer than the corresponding bonds in the saturated entities and the N-terminal amide system is distorted. Ac-ΔAla-NHMe constitutes an exception. ItsC-terminal amide bond is shorter than the standard one and both amide systems are ideally planar. Ac-(E)-ΔAbu-NHMe shares stereoelectronic features with both Ac-ΔAla-NHMe and (Z)-dehydroamides.  相似文献   

14.
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   

15.
The α1 subunit coding for the human brain type E calcium channel (Schneider et al., 1994) was expressed in Xenopus oocytes in the absence, and in combination with auxiliary α2δ and β subunits. α1E channels directed with the expression of Ba2+ whole-cell currents that completely inactivated after a 2-sec membrane pulse. Coexpression of α1E with α2bδ shifted the peak current by +10 mV but had no significant effect on whole-cell current inactivation. Coexpression of α1E with β2a shifted the peak current relationship by −10 mV, and strongly reduced Ba2+ current inactivation. This slower rate of inactivation explains that a sizable fraction (40 ± 10%, n= 8) of the Ba2+ current failed to inactivate completely after a 5-sec prepulse. Coinjection with both the cardiac/brain β2a and the neuronal α2bδ subunits increased by ≈10-fold whole-cell Ba2+ currents although coinjection with either β2a or α2bδ alone failed to significantly increase α1E peak currents. Coexpression with β2a and α2bδ yielded Ba2+ currents with inactivation kinetics similar to the β2a induced currents, indicating that the neuronal α2bδ subunit has little effect on α1E inactivation kinetics. The subunit specificity of the changes in current properties were analyzed for all four β subunit genes. The slower inactivation was unique to α1E2a currents. Coexpression with β1a, β1b, β3, and β4, yielded faster-inactivating Ba2+ currents than currents recorded from the α1E subunit alone. Furthermore, α1E2bδ/β1a; α1E2bδ/β1b; α1E2bδ/β3; α1E2bδ/β4 channels elicited whole-cell currents with steady-state inactivation curves shifted in the hyperpolarized direction. The β subunit-induced changes in the properties of α1E channel were comparable to modulation effects reported for α1C and α1A channels with β3≈β1b > β1a≈β4≫β2a inducing fastest to slowest rate of whole-cell inactivation. Received: 27 March 1997/Revised: 10 July 1997  相似文献   

16.
Systematic screening of single-gene knockout collection of Escherichia coli BW25113 (the Keio collection) was performed to select mutants that could enhance the deethylation of 7-ethoxycoumarin catalyzed by CYP154A1. After 96-well plate high-throughput screening followed by test tube assays, four mutants (ΔcpxA, ΔgcvR, ΔglnL, and an unknown-gene-deleted one (Δuk)) were able to increase the CYP154A1 activity by approximately 1.4–1.7 times compared with that of the control strain. When new mutants were constructed by disrupting individually the cpxA, gcvR, glnL, and uk genes in E. coli BW25113, three of them (ΔcpxA, ΔgcvR, and ΔglnL) showed high levels of CYP154A1 activity. However, the uk-disruptant failed to enhance the CYP154A1 activity, suggesting that the high CYP154A1 activity of the Δuk mutant in the Keio collection was due to a spontaneous mutation in the chromosome. In-frame deletion mutants of ΔcpxA, ΔgcvR, and ΔglnL also exhibited high enzyme activity, and complementation of these mutations could decrease CYP154A1 activity. These results indicated that the enhancement of the enzyme activity was not caused by polar effects on their neighbor genes. To our knowledge, this is the first report on a genome-wide screening of the genes for deletion to improve the activity of a recombinant whole-cell biocatalyst.  相似文献   

17.
Southern blots of genomic DNA from 23 strains of laboratory mice and 19 individual wild mice were examined for restriction fragment length polymorphisms in their loci encoding the T-cell receptors (Tcr): the constant regions of the α, β, and γ chains (C α,C β, andC γ) and a variable region family of the β chain (V β8). Only a few polymorphisms were observed for each locus in the laboratory mice after using three restriction enzymes,Bam HI,Eco RI, andHind III. All the laboratory mice examined fall into one of two types for theC α,C β andV β8 loci and one of three types for theC γ. These types are found in some of the wild mice studied, indicating that they were already present in the founder mice of laboratory mouse strains. In contrast, theTcr genes are highly polymorphic among wild mice. Analysis of the polymorphisms in these loci suggests that laboratory mice have inherited their genes not only fromMus musculus domesticus, but also from other subspecies, and much more than previously believed from Asian subspecies.  相似文献   

18.
 The stearoyl-acyl carrier protein Δ9 desaturase (Δ9D) uses an oxo-bridged diiron center to catalyze the NAD(P)H– and O2–dependent desaturation of stearoyl-ACP. Δ9D, ribonucleotide reductase, and methane monooxygenase have substantial similarities in their amino acid primary sequences and the physical properties of their diiron centers. These three enzymes also appear to share common features of their reaction cycles, including the binding of O2 to the diferrous state and the subsequent generation of transient diferric-peroxo and diferryl species. In order to investigate the coordination environment of the proposed diferric-peroxo intermediate, we have studied the binding of azide to the diiron center of Δ9D using optical, resonance Raman (RR), and transient kinetic spectroscopic methods. The addition of azide results in the appearance of new absorption bands at 325 nm and 440 nm (k app≈3.5 s–1 in 0.7 M NaN3, pH 7.8). RR experiments demonstrate the existence of two different adducts: an η1–terminal structure at pH 7.8 (14N3 asymmetric stretch at 2073 cm–1, resolved into two bands with 15N14N2 ) and a μ-1,3 bridging structure at pH<7 (14N3 asymmetric stretch at 2100 cm–1, shifted as a single band with 15N14N2 ). Both adducts also exhibit an Fe–N3 stretching mode at ≈380 cm–1, but no accompanying Fe–O–Fe stretching mode, presumably due to either protonation or loss of the oxo bridge. The ability to form a μ-1,3 bridging azide supports the likelihood of a μ-1,2 bridging peroxide as a catalytic intermediate in the Δ9D reaction cycle and underscores the adaptability of binuclear sites to different bridging geometries. Received: 23 August 1996 / Accepted: 4 October 1996  相似文献   

19.
 Nucleotide sequences for the three exons of the β2-microglobulin (β2m) gene (B2m) were determined for 135 animals representing 37 species and all 16 genera of neotropical primates (Platyrrhini). Twenty-eight different nucleotide sequences, encoding for 26 different proteins, were obtained. In comparison with those of other primate species, the β2-microglobulins of the Platyrrhini form a distinct clade. Individual genera of neotropical primates have distinctive B2m sequences, but within a genera species can have either the same or different B2m sequences. B2m polymorphism was found within three of the species sampled: Callicebus personatus, Saguinus midas, and Aotus azarae. Of these only the polymorphism in A. azarae has an effect upon the mature, functional β2m protein: residue 4 being either alanine or threonine. The A. azarae B2m allele encoding alanine at position 4 is shared with another species of Aotus (A. infulatus). In pairwise comparison the mature β2m proteins of neotropical primates differ by 1–9 amino acid substitutions which can occur at 18 positions within the sequence. The substitutions are distributed throughout the primary structure but are more commonly found in loops rather than β strands of the tertiary structure. Of 17 residues of β2m which hydrogen-bond with the class I heavy chain in human MHC class I molecules, 13 are conserved in the neotropical primates. The overall pattern of sequence variation in the B2m genes of the Platyrrhini is consistent with an evolution by successive selectively neutral events. Received: 13 November 1997 / Revised: 12 January 1998  相似文献   

20.
β-Glucosidase and β-galactosidase activity profile tested in different seeds during 24 h germination revealed reasonably high levels of activity inVigna radiata, Cicer arietinum, andTrigonella foenum-graecum. In all seeds tested, β-galactosidase activity was, in general, higher than that of β-glucosidase.T. foenum-graecum seedlings exhibited maximal total and specific activities for both the enzymes during 72 h germination. Se supplementation as Na2SeO3 up to 0.75 ppm was found to be beneficial to growth and revealed selective enhancement of β-galactosidase activity by 40% at 0.5 ppm Se. The activities of both the enzymes drastically decreased at 1.0 ppm level of Se supplementation. On the contrary, addition of Na2SeO3 in vitro up to 1 ppm to the enzyme extracts did not influence these activities. Hydrolytic rates of β-glucosidase in both control and Se-supplemented groups were enhanced by 20% with 0.05M glycerol in the medium and 30% at 0.1M glycerol. The rates were marginally higher in Se-supplemented seedlings than the controls, irrespective of added glycerol in the medium. In contrast, hydrolysis by β-galactosidase showed a trend of decrease in Se-supplemented seedlings compared to the control, when glycerol was present in the medium. Addition of Se in vitro in the assay medium showed no difference in the hydrolytic rate by β-galactosidase when compared to control, while the activity of β-glucosidase declined by 50%. Se-grown seedlings showed an enhancement of transglucosidation rate by 40% in the presence of 0.1M glycerol. The study reveals a differential response to Se among the β-galactosidase and β-glucosidase ofT. foenumgraecum with increase in the levels of β-galactosidase activity.  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