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.
Abstract

Adenosine derivatives bearing in 2-position the (R,S)- phenylhydroxypropynyl chain were evaluated for their potency at human A2B adenosine receptor, stably transfected on CHO cells, on the basis that (R,S)-2-phenylhydroxy-propynyl-5′-N-ethylcarboxyamidoadenosine [(R,S)-PHPNECA] was found to be a good agonist at the A2B receptor subtype. Biological studies demonstrated that the presence of small alkyl groups in N 6-position of these molecules are well tolerated, whereas large groups abolished A2B potency. On the other hand, the presence of an ethyl group in the 4′-carboxamido function seems to be optimal, the (S)-PHPNECA resulting the most potent agonist at A2B receptor reported so far.  相似文献   

2.
Summary The reagent sym-triazine trichloride is used as a bifunctional reagent to generate RNA-protein cross-links within intact ribosomal subunits from E. coli. The reaction takes place in a stepwise manner, involving substitution of one chlorine atom at 12° and pH 8, and substitution of the second at 40° and pH 6. The cross-linked proteins are analysed by two-dimensional electrophoresis, and the existence of a stable cross-linkage is demonstrated by isolating protein-oligonucleotide complexes from 32P-labelled subunits. The proteins cross-linked are S3 and S4 in the 30S subunit, and L2 in the large subunit, together with smaller amounts of other proteins. The reagent should prove useful in topographical studies of the E. coli ribosome as it is a rigid molecule and generates very short cross-links.  相似文献   

3.
The Rps0 proteins of Saccharomyces cerevisiae are components of the 40S ribosomal subunit required for maturation of the 3′ end of 18S rRNA. Drosophila and human homologs of the Rps0 proteins physically interact with Rps21 proteins, and decreased expression of both proteins in Drosophila impairs control of cellular proliferation in hematopoietic organs during larval development. Here, we characterize the yeast RPS21A/B genes and show that strains where both genes are disrupted are not viable. Relative to the wild type, cells with disrupted RPS21A or RPS21B genes exhibit a reduction in growth rate, a decrease in free 40S subunits, an increase in the amount of free 60S subunits, and a decrease in polysome size. Ribosomal RNA processing studies reveal RPS21 and RPS0 mutants have virtually identical processing defects. The pattern of processing defects observed in RPS0 and RPS21 mutants is not a general characteristic of strains with suboptimal levels of small subunit ribosomal proteins, since disruption of the RPS18A or RPS18B genes results in related but distinct processing defects. Together, these data link the Rps0 and Rps21 proteins together functionally in promoting maturation of the 3′ end of 18S rRNA and formation of active 40S ribosomal subunits.  相似文献   

4.
The oligomeric form of the larger subunit designated as Am produced by alkali treatment of ribulose-1,5-diphosphate carboxylase from the purple sulfur bacterium, Chromatium strain D, retained partial enzymic activity in the absence of the small subunit (B). Supporting evidence was obtained by polyacrylamide gel electrophoresis at pH 8.9 and Sephadex G-200 gel filtration equilibrated with alkaline buffer at pH 9.2. The specific enzyme activity of Am (45 nmoles CO2 fixed/mg protein/min) was approximately 15% of the native intact enzyme molecule. By sodium dodecyl sulfatepolyacrylamide gel electrophoresis, the Am preparation was proved to be free from contamination of subunit B. With reservation of the sensitivity limit of this particular technique we concur that the larger subunit is the catalytic entity of the carboxylase reaction. The optimum pH of the ribulose-1,5-diphosphate carboxylase reaction catalyzed by isolated Am lies on the alkaline side at about pH 8.3 with or without Mg2+. The undissociated native enzyme possesses an optimum pH on the alkaline side in the absence of Mg2+, which shifts to the acidic side in the presence of Mg2+. From this behavior it is inferred that the association of the smaller subunit with the larger subunit causes conformational stabilization of the enzyme molecule with an accompanying change in the pH optimum due to Mg2+.  相似文献   

5.
Soybean mutant lines that differ in 11S glycinin and 7S β-conglycinin seed storage protein subunit compositions were developed. These proteins have significant influence on tofu quality. The molecular mechanisms underlying the mutant lines are unknown. In this study, gene-specific markers for five of the glycinin genes (Gy1 to Gy5) were developed using three 11S null lines, two A4 null Japanese cultivars, Enrei and Raiden, and a control cultivar, Harovinton. Whereas gene-specific primers produced the appropriate products in the control cultivar for the Gy1, Gy2, Gy3 and Gy5 genes, they did not amplify in mutants missing the A1aB2, A2B1a, A1b B1b, and A3B4 subunits. However, ecotype targeting induced local lesions in genomes (EcoTILLING) and sequencing analysis revealed that the absence of the A4 peptide in the mutants is due to the same point mutation as that in Enrei and Raiden. Selection efficiency of the gene-specific primer pairs was tested using a number of breeding lines segregating for the different subunits. Primer pairs specific to each of the Gy1, Gy2, Gy3, and Gy5 genes can be used to detect the presence or absence of amplification in normal or mutant lines. The Gy4 null allele can be selected for by temperature-switch PCR (TS-PCR) for identification of the A4 (G4) null genotypes. In comparison to protein analysis by SDS-PAGE, gene-specific markers are easier, faster and more accurate for analysis, they do not have to use seed, and can be analyzed at any plant growth stage for marker-assisted selection.  相似文献   

6.
Summary Antibodies to individual chloroplast ribosomal (r-)proteins ofChlamydomonas reinhardtii synthesized in either the chloroplast or the cytoplasm were used to examine the relatedness ofChlamydomonas r-proteins to r-proteins from the spinach (Spinacia oleracea) chloroplast,Escherichia coli, and the cyanobacteriumAnabaena 7120. In addition,35S-labeled chloroplast r-proteins from large and small subunits ofC. reinhardtii were coelectrophoresed on 2-D gels with unlabeled r-proteins from similar subunits of spinach chloroplasts,E. coli, andAnabaena to compare their size and net charge. Comigrating protein pairs were not always immunologically related, whereas immunologically related r-protein pairs often did not comigrate but differed only slightly in charge and molecular weight. In constrast, when35S-labeled chloroplast r-proteins from large and small subunits of a closely related speciesC. smithii were coelectrophoresed with unlabeledC. reinhardtii chloroplast r-proteins, only one pair of proteins from each subunit showed a net displacement in mobility.Analysis of immunoblots of one-dimensional SDS and two-dimensional urea/SDS gels of large and small subunit r-proteins from these species revealed more antigenic conservation among the four species of large subunit r-proteins than small subunit r-proteins.Anabaena r-proteins showed the greatest immunological similarity toC. reinhardtii chloroplast r-proteins. In general, antisera made against chloroplast-synthesized r-proteins inC. reinhardtii showed much higher levels of cross-reactivity with r-proteins fromAnabaena, spinach, andE. coli than did antisera to cytoplasmically synthesized r-proteins. All spinach r-proteins that cross-reacted with antisera to chloroplast-synthesized r-proteins ofC. reinhardtii are known to be made in the chloroplast (Dorne et al. 1984b). FourE. coli r-proteins encoded by the S10 operon (L2, S3, L16, and L23) were found to be conserved immunologically among the four species. Two of the large subunit r-proteins, L2 and L16, are essential for peptidyltransferase activity. The third (L23) and two otherE. coli large subunit r-proteins (L5 and L27) that have immunological equivalents among the four species are functionally related to but not essential for peptidyltransferase activity.  相似文献   

7.
Six tyrosinase isozymes were purified from the browned gill of the fruiting body of Lentinus edodes by ammonium sulfate fractionation, DEAE-Sephacel and Q-Sepharose column chromatography, and partially denaturing SDS–PAGE. At the step of Q-Sepharose column chromatography, two active fractions (A and B) were obtained. Each fraction was separated to three further fractions, A1, A2, and A3, and B1, B2, and B3, respectively, by partially denaturing SDS–PAGE. All these isozymes consisted of two types of polypeptides: a polypeptide (Aα or Bα) and either β (Aβ or Bβ) or γ polypeptide (Aγ or Bγ). The α polypeptide contained the consensus amino acid sequence of the active site of known tyrosinases, which is considered to act as a catalytic subunit. From the results of peptide mapping and the amino acid composition, Aα and Bα polypeptides were considered to be different proteins. The kinetic properties of the purified tyrosinase isozymes differed greatly according to whether they contained β or γ polypeptide, indicating these polypeptides to be a possible regulatory subunit.  相似文献   

8.
1. The objective was to identify the factors driving spatial and temporal variation in annual production (PA) and turnover (production/biomass) ratio (P/BA) of resident brown trout Salmo trutta in tributaries of the Rio Esva (Cantabrian Mountains, Asturias, north‐western Spain). We examined annual production (total production of all age‐classes over a year) (PA) and turnover (P/BA) ratios, in relation to year‐class production (production over the entire life time of a year‐class) (PT) and turnover (P/BT) ratio, over 14 years at a total of 12 sites along the length of four contrasting tributaries. In addition, we explored whether the importance of recruitment and site depth for spatial and temporal variations in year‐class production (PT), elucidated in previous studies, extends to annual production. 2. Large spatial (among sites) and temporal (among years) variation in annual production (range 1.9–40.3 g m?2 per year) and P/BA ratio (range 0.76–2.4 per year) typified these populations, values reported here including all the variation reported globally for salmonids streams inhabited by one or several species. 3. Despite substantial differences among streams and sites in all production attributes, when all data were pooled, annual (PA) and year‐class production (PT) and annual (P/BA) and year‐class P/BT ratios were tightly linked. Annual (PA) and year‐class production (PT) were similar but not identical, i.e. PT = 0.94 PA, whereas the P/BT ratios were 4 + P/BA ratios. 4. Recruitment (Rc) and mean annual density (NA) were major density‐dependent drivers of production and their relationships were described by simple mathematical models. While year‐class production (PT) was determined (R2 = 70.1%) by recruitment (Rc), annual production (PA) was determined (R2 = 60.3%) by mean annual density (NA). In turn, variation in recruitment explained R2 = 55.2% of variation in year‐class P/BT ratios, the latter attaining an asymptote at P/BT = 6 at progressively higher levels of recruitment. Similarly, variations in mean annual density (NA) explained R2 = 52.1% of variation in annual P/BA, the latter reaching an asymptote at P/BA = 2.1. This explained why P/BT is equal to P/BA plus the number of year‐classes at high but not at low densities. 5. Site depth was a major determinant of spatial (among sites) variation in production attributes. All these attributes described two‐phase trajectories with site depth, reaching a maximum at sites of intermediate depth and declining at shallower and deeper sites. As a consequence, at sites where recruitment and mean annual density reached minimum or maximum values, annual (PA) and year‐class production (PT) and annual (P/BA) and year‐class P/BT ratios also reached minimum and maximum values.  相似文献   

9.
clpC ofBacillus subtilis is part of an operon containing six genes. Northern blot analysis suggested that all genes are co-transcribed and encode stress-inducible proteins. Two promoters (PA and PB) were mapped upstream of the first gene. PA resembles promoters recognized by the vegetative RNA polymerase EσA. The other promoter (PB) was shown to be dependent on σB, the general stress σ factor in B. subtilis, suggesting that clpC, a potential chaperone, is expressed in a σB-dependent manner. This is the first evidence that σB in B, subtilis is involved in controlling the expression of a gene whose counterpart, clpB, is subject to regulation by σ32 in Escherichia coli, indicating a new function of σB-dependent general stress proteins. PB deviated from the consensus sequence of σB promoters and was only slightly induced by starvation conditions. Nevertheless, strong induction by heat, ethanol, and salt stress occurred at the σB-dependent promoter, whereas the vegetative promoter was only weakly induced under these conditions. However, in a sigB mutant, the σA-like promoter became inducible by heat and ethanol stress, completely compensating for sigB deficiency. Only the downstream σA-like promoter was induced by certain stress conditions such as hydrogen peroxide or puromycin. These results suggest that novel stress-induction mechanisms are acting at a vegetative promoter. Involvement of additional elements in this mode of induction are discussed.  相似文献   

10.
The pH dependence of emission peak temperature and decay time of thermoluminescence arising from S2QB and S2QA recombinations demonstrates that a stabilization of S2QB occurs at low pH whereas stabilization of S2QA occurs at high pH. Based on comparative analysis of thermoluminescence parameters of the two types of recombination, we suggest that in the pH range between 5.3 and 7.5, Em(S2/S1) and Em(QA/QA ) are constant, but Em(QB/QB ) gradually increases with decreasing pH, while in the pH range between 7.5 and 8.5, an unusual change occurs on S2QA charge pair, which is interpreted as either a decrease in Em(S2/S1) or an increase in Em(QA/QA ).  相似文献   

11.
The effects of various metabolites on the two most common phosphoglucomutase allozymes (PGMA and PGMB) in Drosophila melanogaster have been investigated in vitro. 2,3-Diphosphoglycerate (2,3DPG) inhibited PGMA and PGMB to the same degree in the presence of 25 µM glucose-1,6-diphosphate (G1, 6P2). However a higher concentration of G1,6P2 partially reversed the inhibition of PGMA exerted by 2,3DPG, so that in the presence of 150 µM G1,6P2 the inhibition of PGMA was half that of PGMB at pH 6.0. Glycerol-3-phosphate (G3P) had no significant effect at pH 7.4 but exerted an activating effect at pH 6.0 which was more pronounced in the case of PGMB. ATP, citrate, and fructose-1, 6-diphosphate (F1,6P2) inhibited both PGMA and PGMB. The differences found in vitro between these two allozymes can have a significant impact on in vivo function and, therefore, on the maintenance of PGM polymorphism in experimental populations of D. melanogaster studied in the laboratory.  相似文献   

12.
Summary Two new forms of the plasma membrane ATP-ase ofMicrococcus lysodeikticus NCTC 2665 were isolated from a sub-strain of the microorganism by polyacrylamide gel electrophoresis. One of them had a mol.wt of 368,000 and a very low specific activity (0.80 µ mol.min–1.mg protein–1) that could not be stimulated by trypsin. This form has been called BI (strain B, inactive). If the electrophoresis was carried out in the presence of reducing agents (i.e., dithiothreitol) and the pH of the effluent maintained at a value of 8.5 another form of the enzyme was obtained. This had a mol.wt of 385,000 and a specific activity of 2.5–5.0 µ mol.min–1.mg protein–1 that could be stimulated by trypsin to 5–10 µ mol.min–1.mg protein–1. This preparation of the ATPase has been called form BA (strain B, enzyme active). The subunit composition of both forms has been studied by sodium dodecyl sulphate and urea gel electrophoresis and compared to that of the enzyme previously purified from the original strain (form A). The three forms of the enzyme had similar and subunits, with mol.wt of about 50,000 and 30,000 dalton, respectively. They also had in common the component(s) of relative mobility 1.0, whose status as true subunit(s) of the enzyme remains yet to be established. However, subunit, that had a mol.wt of about a 52,500 in form A (Andreu et al. Eur. J. Biochem. (1973) 37, 505–515), had a mol.wt similar to in form BI and about 60,000 in form BA. Furthermore BA usually showed two types of this subunit ( and) and an additional peptide chain () with a mol.wt of about 25,000 dalton. This latter subunit seemed to account for the stimulation by trypsin of form BA.Forms BA could be converted to BI by storage and freezing and thawing. Conventional protease activity could not be detected in any of the purified ATPase forms and addition of protease inhibitors to form BA failed to prevent its conversion to form BI. The low activity form (BI) was more stable than the active forms of the enzyme and also differed in its circular dichroism. These results show thatM. lysodeikticus ATPase can be isolated in several forms. Although these variations may be artifacts caused by the purification procedures, they provide model systems for understanding the structural and functional relationships of the enzyme and for drawing some speculations about its functionin vivo.  相似文献   

13.
Summary The area-specific coductance of the membrane in the acid and basic zones (denoted byG A andG B , respectively) ofChara cells was measured in flowing solutions, containing 5mm zwitterionic buffer, as a function of the external pH(denoted by pH0). During illuminationG A was 1 S/m2 for pH0 in the range 5 to 8.5, and increased markedly to 3 to 6 S/m2 at higher pH0.G B , however, was always larger thanG A during illumination with a typical magnitude of 5 to 15 S/m2 for pH0 6 to 12. Thus under many experimental conditions it is possible that there is no single correct value for the membrane area-specific conductance. A flow of current in the external medium between the acid and basic regions was found to be associated with pH banding, and also withG B exceedingG A . This current could be present in flowing solutions without added HCO 3 over a wide range of pH0 and at high (25mm) buffer concentration. Combining measurements ofG A andG B with measurements of the currents in the acid and basic zones (denoted byJ A andJ B , respectively), it was estimated that the resting (i.e. in the absence of net current flow) potential difference (PD) across the membranes within the individual zones (denoted byU A andU B ) was –265±20 and –183±5 mV, respectively, during illumination. Upon the removal of illumination at pH0-7.5,G A ,G B andJ B were found to decrease rapidly during the initial few hundred seconds. During this period (U B V m ) remained relatively constant. A transient hyperpolarization ofV m often occurred, the magnitude of which was correlated with the magnitude ofJ B prior to the removal of illumination. After some 0.5 to 1 ksec of darkness,G A andG B had both decreased considerably and nowG A G B andU A U B V m . Eventually, after 2 to 8 ksec of darkness, the membrane conductance was effectively homogeneous with a much smaller magnitude (typically<0.2S/m2) andV m was depolarized by typically 5 to 15 mV.  相似文献   

14.
The loci in the major histocompatibility complex (MHC) of the rat which code for class I and class II antigens—RT1.A and RT1.B, respectively — have previously been separated by laboratory-derived recombinants and by observations in inbred and wild rats. Closely linked to the MHC is the growth and reproduction complex (Grc) which contains genes influencing body size (dw-3) and fertility (ft). These phenotypic markers were used in this study to orient the A and B loci of the MHC. Two recombinants were used for mapping. The BIL(R1) animal is a recombinant between the MHC and Grc, and it carries the haplotype RT1.A lBlGrc+. The r10 animal is an intra-MHC recombinant, and it has the haplotype RT1.A nB1 Grc. These recombinants were characterized serologically, by mixed lymphocyte reactivity, by immune responsiveness to poly (Glu52Lys33Tyr15) and by the presence of the dw-3 gene. The data demonstrate that the gene order of the loci is: dw-3-RT1.B-RT1.A.  相似文献   

15.
Growth of Delftia acidovorans MC1 on 2,4-dichlorophenoxyacetic acid (2,4-D) and on racemic 2-(2,4-dichlorophenoxy)propanoic acid ((RS)-2,4-DP) was studied in the perspective of an extension of the strain’s degradation capacity at alkaline pH. At pH 6.8 the strain grew on 2,4-D at a maximum rate (μmax) of 0.158 h−1. The half-maximum rate-associated substrate concentration (Ks) was 45 μM. At pH 8.5 μmax was only 0.05 h−1 and the substrate affinity was mucher lower than at pH 6.8. The initial attack of 2,4-D was not the limiting step at pH 8.5 as was seen from high dioxygenase activity in cells grown at this pH. High stationary 2,4-D concentrations and the fact that μmax with dichlorprop was around 0.2 h−1 at both pHs rather pointed at limited 2,4-D uptake at pH 8.5. Introduction of tfdK from D. acidovorans P4a by conjugation, coding for a 2,4-D-specific transporter resulted in improved growth on 2,4-D at pH 8.5 with μmax of 0.147 h−1 and Ks of 267 μM. Experiments with labeled substrates showed significantly enhanced 2,4-D uptake by the transconjugant TK62. This is taken as an indication of expression of the tfdK gene and proper function of the transporter. The uncoupler carbonylcyanide m-chlorophenylhydrazone (CCCP) reduced the influx of 2,4-D. At a concentration of 195 μM 2,4-D, the effect amounted to 90% and 50%, respectively, with TK62 and MC1. Cloning of tfdK also improved the utilization of 2,4-D in the presence of (RS)−2,4-DP. Simultaneous and almost complete degradation of both compounds occurred in TK62 up to D = 0.23 h−1 at pH 6.8 and up to D = 0.2 h−1 at pH 8.5. In contrast, MC1 left 2,4-D largely unutilized even at low dilution rates when growing on herbicide mixtures at pH 8.5.  相似文献   

16.
To understand how environmental changes have influenced forest productivity, stemwood biomass (B) dynamics were analyzed at 1267 permanent inventory plots, covering a combined 209 ha area of unmanaged temperate‐maritime forest in southwest British Columbia, Canada. Net stemwood production (ΔB) was derived from periodic remeasurements of B collected over a 40‐year measurement period (1959–1998) in stands ranging from 20 to 150 years old. Comparison between the integrated age response of net stemwood production, ΔB(A), and the age response of stemwood biomass, B(A), suggested a 58 ± 11% increase in ΔB between the first 40 years of the chronosequence period (1859–1898) and the measurement period. To estimate extrinsic forcing on ΔB, several different candidate models were developed to remove variation explained by intrinsic factors. All models exhibited temporal bias, with positive trends in (observed minus predicted) residual ΔB ranging between of 0.40 and 0.64% yr?1. Applying the same methods to stemwood growth (G) indicated residual increases ranging from 0.43 and 0.67% yr?1. Higher trend estimates corresponded with models that included site index (SI) as a predictor, which may reflect exaggeration of the age‐decline in SI tables. Choosing a model that excluded SI, suggested that ΔB increased by 0.40 ± 0.18% yr?1, while G increased by 0.43 ± 0.12% yr?1 over the measurement period. Residual G was significantly correlated with atmospheric carbon dioxide (CO2), temperature (T), and climate moisture index (CMI). However, models driven with climate and CO2, alone, could not simultaneously explain long‐term and measurement‐period trends without additional representation of indirect effects, perhaps reflecting compound interest on direct physiological responses to environmental change. Evidence of accelerating forest regrowth highlights the value of permanent inventories to detect and understand systematic changes in forest productivity caused by environmental change.  相似文献   

17.
The ferriheme resonances of the low-spin (S = 1/2) complexes of wild-type (wt) nitrophorin 2 (NP2) and its heme pocket mutant NP2(V24E) with imidazole (ImH), histamine (Hm), and cyanide (CN) as the sixth ligand have been investigated by NMR spectroscopy as a function of pH (4.0–7.5). For the three wt NP2 complexes, the ratio of the two possible heme orientational isomers, A and B, remains almost unchanged (ratio of A:B approximately 1:6 to 1:5) over this wide pH range. However, strong chemical exchange cross peaks appear in the nuclear Overhauser effect spectroscopy/exchange spectroscopy (NOESY/EXSY) spectra for the heme methyl resonances at low pH (pH* 4.0–5.5), which indicate chemical exchange between two species. We have shown these to be two different exogenous ImH or Hm orientations that are denoted B and B′, with the ImH plane nearly parallel and perpendicular to the ImH plane of the protein-provided His57, respectively. The wt NP2–CN complex also shows EXSY cross peaks due to chemical exchange, which is shown to be a result of interchange between two ruffling distortions of the heme. The same ruffling distortion interchange is also responsible for the ImH and Hm chemical exchange. For the three NP2(V24E) ligand complexes, no EXSY cross peaks are observed, but the A:B ratios change dramatically with pH. The fact that heme favors the A orientation highly for NP2(V24E) at low pH as compared with wt NP2 is believed to be due to the steric effect of the V24E mutation. The existence of the B′ species at lower pH for wt NP2 complexes and the increase in A heme orientation at lower pH for NP2(V24E) are believed to be a result of a change in structure near Glu53 when it is protonated at low pH. 1H{13C} heteronuclear multiple quantum coherence (HMQC) spectra are very helpful for the assignment of heme and nearby protein side chain resonances.  相似文献   

18.
A and FB. The g-tensor orientation of FA and FB is believed to be correlated to the preferential localization of the mixed-valence and equal-valence (ferrous) iron pairs in each [4Fe-4S]+ cluster. The preferential position of the mixed-valence and equal-valence pairs, in turn, can be inferred from the study of the temperature dependence of contact-shifted resonances by 1H NMR spectroscopy. For this, a sequence-specific assignment of these signals is required. The 1H NMR spectrum of reduced, unbound PsaC from Synechococcus sp. PCC 7002 at 280.4 K in 99% D2O solution shows 18 hyperfine-shifted resonances. The non-solvent-exchangeable, hyperfine-shifted resonances of reduced PsaC are clearly identified as belonging to the cysteines coordinating the clusters FA and FB by their downfield chemical shifts, by their temperature dependencies, and by their short T 1 relaxation times. The usual fast method of assigning the 1H NMR spectra of reduced [4Fe-4S] proteins through magnetization transfer from the oxidized to the reduced state was not feasible in the case of reduced PsaC. Therefore, a de novo self-consistent sequence-specific assignment of the hyperfine-shifted resonances was obtained based on dipolar connectivities from 1D NOE difference spectra and on longitudinal relaxation times using the X-ray structure of Clostridium acidi urici 2[4Fe-4S] cluster ferredoxin at 0.94 Å resolution as a model. The results clearly show the same sequence-specific distribution of Curie and anti-Curie cysteines for unbound, reduced PsaC as established for other [4Fe-4S]-containing proteins; therefore, the mixed-valence and equal-valence (ferrous) Fe-Fe pairs in FA and FB have the same preferential positions relative to the protein. The analysis reveals that the magnetic properties of the two [4Fe-4S] clusters are essentially indistinguishable in unbound PsaC, in contrast to the PsaC that is bound as a component of the PS I complex. Received: 1 February 2000 / Accepted: 20 March 2000  相似文献   

19.
For the usual full rank univariate least squares regression model y = XB + e, E(e) = 0, E(ee) = A, the equality of the estimates occurs when B-B* = (XA?1X)?1XA-1y-(XX)?1Xy = 0. A necessary and sufficient condition for this equality is that A has some N - k + 1 roots equal where N is the rank of A and k is the rank of X.  相似文献   

20.
15N-enriched (D ,L -Leu)n, (γ-OMe-D ,L -Glu)n, (D ,L -Val)n, and (D ,L -Phe)n were prepared, 40.55-MHz 15N-nmr spectra were measured in various solvents. The signal patterns depend strongly on the nature of the solvent, yet in most cases at least four signals are resolved, representing the four enantiomeric pairs of triads L -L -L (D -D -D ), L -D -L (D -L -D ), L -L -D (D -D -L ), and D -L -L (L -D -D ). Numerous copolypeptides of the general structure (A)n-B*-(A)m (the asterisk denotes 40–50% 15N enrichment) were synthesized and measured as models for syndiotactic sequences in the spectra of poly(D ,L -amino acids). In this way unambiguous assignments for both isotactic and syndiotactic trials were obtained. A spectroscopic rule was established: “isotactic sequences absorb downfield of syndiotactic ones.” Furthermore, the spectra of various types of stereocopolypeptides such as (L -Leu/L -Val)n and (L -Leu/D -Val)n were investigated, including the ternary systems (L -Leu/L -Ala/D -Ala)n (L -Leu/L -Ala/Gly)n, (L -Leu/D -Ala/Gly)n, (L -Val/L -Ala/Gly)n, and (L -Val/D -Ala/Gly)n. All copolymerization of D - and L -amino acid NCAs investigated in this work showed a low degree of stereoselectivity.  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