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.
Like other chenopods, sugarbeets (Beta vulgaris L. cv Great Western D-2) accumulate glycine betaine when salinized; this may be an adaptive response to stress. The pathway of betaine synthesis in leaves of salinized (150-200 millimolar NaCl) sugarbeet plants was investigated by supplying [14C]formate, phosphoryl[14C]monomethylethanolamine ([14C][unk] MME) or phosphoryl[14C]choline ([14C][unk] choline) to leaf discs and following 14C incorporation into prospective intermediates. The 14C kinetic data were used to develop a computer model of the betaine pathway.

When [14C]formate was fed, [unk] MME, phosphoryldimethylethanolamine ([unk] DME) and [unk] choline were the most prominent methylated products at short labeling times, after which 14C appeared in free choline and in betaine. Phosphatidylcholine labeled more slowly than [unk] choline, choline, and betaine, and behaved as a minor end product. Very little 14C entered the free methylethanolamines. When [14C][unk] MME was supplied, a small amount was hydrolyzed to the free base but the major fate was conversion to [unk] DME, [unk] choline, free choline, and betaine; label also accumulated slowly in phosphatidylcholine. Label from supplied [14C][unk] choline entered choline and betaine rapidly, while phosphatidylcholine labeled only slowly and to a small extent.

These results are consistent with the pathway [unk] MME →[unk] DME → [unk] choline → choline → → betaine, with a minor side branch leading from [unk] choline into phosphatidylcholine. This contrasts markedly (a) with the pathway of stress-induced choline and betaine synthesis in barley, in which phosphatidylcholine apparently acts as an intermediate (Hitz, Rhodes, Hanson 1981, Plant Physiol 68: 814-822); (b) with choline biogenesis in mammalian liver and microorganisms. Computer modeling of the experimental data pointed strongly to regulation at the [unk] choline → choline step, and also indicated that the rate of [unk] choline synthesis is subject to feedback inhibition by [unk] choline.

  相似文献   

2.
Conformations of 8-bromo-2′-[unk]-triisopropylbenzenesulfonyladenosine ([unk]) and its 3′-[unk]-isomer ([unk]) in solution have been determined by the use of intramolecular nuclear Overhauser effects in 1H NMR spectroscopy. Compound [unk] has been proved to have a conformation in which the adenosine and benzene rings are intramolecularly stacked and compound [unk] an elongated non-stacked conformation in dimethylsulphoxide. The 5′-[unk]-acetyl derivative of [unk] has also been found to adopt the intramolecularly stacked conformation in dimethylsulphoxide, but a non-stacked one in chloroform. Coupling constants observed are discussed in connection with the conformation of the ribose moiety. The 13C NMR spectra have also been examined, but no effect which could be ascribed to the stacking phenomena was observed in the carbon chemical shifts.  相似文献   

3.
Wheat plants (Triticum aestivum L.) were grown at the same photosynthetic photon flux (PPF), 200 micromoles per square meter per second, but with phytochrome photoequilibrium ([unk]) values of 0.81, 0.55, and 0.33. Plants grown at [unk] values of 0.55 and 0.33 tillered 43 and 56%, less compared with plants grown at [unk] of 0.81. Main culm development (Haun stage) was slightly more advanced at lower values of [unk], and leaf sheaths, but not leaf lamina, were longer at lower [unk]. Dry-mass accumulation was not affected by different levels of [unk]. Three levels of PPF (100, 200, and 400 micromoles per square meter per second) and two lamp types, metal halide and high pressure sodium, were also tested. Higher levels of PPF resulted in more dry mass, more tillering, and a more advanced Haun stage. There was no difference in plant dry mass or development under metal halide versus high pressure sodium lamps, except for total leaf length, which was greater under high pressure sodium lamps (49.5 versus 44.9 centimeters, P < 0.01).  相似文献   

4.
Mn2+, Cu2+, and nitroxyl amines have been shown to bond to plant homopolygalacturonan matrices in a spatially sequential fashion. As a consequence of this special form of cooperativity the lattice constant (κ), determined from Van Vleck's second moment relationship, approaches 1 only when the average number of dipolar interactions per spin approaches 1 (e.g., an array of dimers). Assuming that one paramagnetic ion or nitroxyl amide pair is bonded per polymer block within the matrix when κ = 1, the anionic ligand's average degree of polymerization ([unk]) can be estimated from the concentration of bonded paramagnetic dimers (e.g., [1/χ]κ~1 = [unk]; χ is the mole fraction of bonded paramagnetic dimers). We have utilized this technique to estimate the average molecular size of homopolygalacturonan blocks in intact higher plant cortical cell walls ([unk] ~83), Nitella cell walls ([unk] ~27) and a commercially available galacturonic acid polymer ([unk] ~35). The [unk] determined from both the intact cortical cell wall lattice and the polygalacturonan were similar to literature values; these findings argue that the electron paramagnetic resonance, (EPR) dipolar spin-spin interaction technique reported herein is a valid approach for estimating molecular size in plant cell walls.  相似文献   

5.
1. The amino acid compositions of human fibrinogen and three intermediate anticoagulant derivatives were determined by column chromatography. The derivatives were isolated by ammonium sulphate fractionation and column electrophoresis from solutions of fibrinogen undergoing spontaneous breakdown. One derivative, isolated as the large electrophoretic peak at the end of the clottable period (100% CP) of the parent fibrinogen solution, was labelled LP(100) and others obtained at twice this period (200% CP) were designated as LP(200) and SP(200) (LP, large peak; SP, small peak). 2. Maximal ;molecular' weights of approx. 294000 for LP(100), 137000 for LP(200) and 37000 for SP(200) were calculated for the protein moieties. At least 265 amino acid residues must have been lost from each fibrinogen molecule during the formation of LP(100), and 1362 during the formation of the other two derivatives. 3. Only one derivative (LP(200)) had a partial specific volume ([unk] 0.725ml./g.) different from that of fibrinogen ([unk] 0.721ml./g.). 4. No significant differences in refractive index at 589mmu were detected. 5. Calculation of the total number of ionizable groups/10(5)g. of each protein moiety showed a preponderance of the following numbers of negative charges: 22 in fibrinogen; 24 in LP(100); 26 in LP(200); 49 in SP(200). The isoionic points were estimated to be approx.+0.03pH unit (for fibrinogen), -0.06pH unit for (LP(100)) and +0.28pH unit (for LP(200)) from the pK of imidazole, and 0.78pH unit above the average pK of aspartyl and glutamyl ions (for SP(200)). These figures agree closely with experimentally determined values of the isoelectric point of fibrinogen and its derivatives.  相似文献   

6.
Inhibitors of caeruloplasmin   总被引:4,自引:3,他引:1  
1. A method is described by which substances inhibiting caeruloplasmin oxidase activity directly may be distinguished from those acting on stimulatory contaminant iron or on the product of enzyme action. 2. Many previously reported inhibitors, including saturated aliphatic carboxylates, hydrazines, 1,10-phenanthroline, borate and various psycho-active drugs, are found either not to act on the enzyme or to inhibit it only weakly. 3. A series of inorganic anions are compared as inhibitors. Anions such as azide and cyanide with strong copper-binding properties are the most effective inhibitors. There is a general inverse relationship between anion size and inhibitory power. Iodide is anomalous, the order of effectiveness of halides being F(-)>I(-)[unk]Cl(-)>Br(-). 4. Multidentate copperchelating ligands have little inhibitory effect. 5. A group of substances containing the structural unit [unk]C=[unk].CO(2)H, including fumarate and benzoate, cause inhibition. 6. Relative inhibitions by a series of mono-substituted benzoates are inversely related to molecular size. 7. Results are discussed in relation to earlier work on the disposition and function of the copper atoms of caeruloplasmin.  相似文献   

7.
Cu(II) affects the yield of cyclobutyl dimers induced in DNA by 254 nm radiation. The effects are a function of r, the ratio of Cu(II) to DNA phosphate, and of the ultraviolet (UV) fluence; they seem to reflect two types of copper complexes with DNA. The first probably involves “exterior” binding to the bases of native DNA and increases [unk]TT formation (without affecting [unk]UT yield) by raising the energy levels of bases other than thymine. The second seems to occur only at high ratios (rs) and only after the structure has been opened locally by UV radiation; it involves “interior” binding of Cu(II) to the bases. This complex tends to decrease dimer yield by holding the bases apart and/or by lowering the energy levels of bases other than thymine. These results illustrate the potential use of DNA photoproducts and ligands to probe the structure and interactions of DNA in vitro and perhaps also in vivo.  相似文献   

8.
The accessibility of the s4U base in native tRNAVal from E.coli was monitored by studying the binding of various mercurials. The relative binding order HgBr2[unk]HgCl2CH3HgOAc[unk]CH3HgCl[unk]PCMB parallels approximately the steric requirements of linear HgX2 or RHgX compounds for SN2 displacement by sulfur, although other factors are operative. Para-chloromercuri-benzoate (PCMB) does not bind the thiolated nucleotide unless the tertiary structure of the tRNA is opened up by removal of Mg2+ ions and heating to 40°. Under these conditions, equilibrium dialysis measurements using 14C-labeled PCMB showed one binding site (n = 0.93) with an association constant, K1, of 9 × 104M−1.  相似文献   

9.
10.
The secondary structure of the isolated tRNA-like sequence (n=159) present at the 3' OH terminus of turnip yellow mosaic virus RNA has been established from partial nuclease digestion with S1 nuclease and T1, CL3, and Naja oxiana RNases. The fragment folds into a 6-armed structure with two main domains. The first domain, of loose structure and nearest the 5' OH terminus, is composed of one large arm which extends into the coat protein cistron. The second, more compact domain, is composed of the five other arms and most probably contains the structure recognized by valyl-tRNA synthetase. In this domain three successive arms strikingly resemble the T[unk], anticodon, and D arms found in tRNA. Near the amino-acid accepting terminus, however, there is a new stem and loop region not found in standard tRNA. This secondary structure is compatible with a L-shaped three-dimensional organization in which the corner of the L and the anticodon-containing limb are similar to, and the amino-acid accepting region different from, that in tRNA. Ethylnitrosourea accessibility studies have shown similar tertiary structure features in the T[unk] loop of tRNAVal and in the homologous region of the viral RNA.  相似文献   

11.
The activity of a 7.3S-8.3S Drosophila DNA polymerase was characterized in detail using poly dA.p(dT)[unk] and poly rA.p(dT)[unk]. With poly dA.p(dT)[unk], Mg(2+) ion was the preferred divalent cation, and enzyme activity was inhibited by K(+) ion and by spermidine. With poly rA.p(dT)[unk], Mn(2+) ion was the preferred divalent cation and enzyme activity was stimulated by K(+) ion and by spermidine. The dependence of enzyme activity on the concentration of primer-template and on the ratio of primer to template was the same in both reactions. The two enzyme activities were identically inhibited by N-ethylmaleimide. Poly dA was replicated extensively and poly rA was replicated partially. The activation energy for poly dA replication was twice that for poly rA replication. Enzyme activity with poly dA.p(dT)[unk] was more stable to thermal inactivation than was enzyme activity with poly rA.p(dT)[unk]. These studies suggest that the same enzyme responds to both the deoxy- and the ribohomopolymer template but that the mechanisms of replication may be different.  相似文献   

12.
The recent proposal of Tipton and Thowsen (Plant Physiol 79: 432-435) that iron-deficient plants reduce ferric chelates in cell walls by a system dependent on the leakage of malate from root cells was tested. Results are presented showing that this mechanism could not be responsible for the high rates of ferric reduction shown by roots of iron-deficient bean (Phaseolus vulgaris L. var Prélude) plants. The role of O2 in the reduction of ferric chelates by roots of iron-deficient bean plants was also tested. The rate of Fe(III) reduction was the same in the presence and in the absence of O2. However, in the presence of O2 the reaction was partially inhibited by superoxide dismutase (SOD), which indicates a role for the superoxide radical, O2[unk], as a facultative intermediate electron carrier. The inhibition by SOD increased with substrate pH and with decrease in concentration of the ferrous scavenger bathophenanthroline-disulfonate. The results are consistent with a mechanism for transmembrane electron transport in which a flavin or quinone is the final electron carrier in the plasma membrane. The results are discussed in relation to the ecological importance that O2[unk] may have in the acquisition of ferric iron by dicotyledonous plants.  相似文献   

13.
The orientational relaxation of the magnetotactic bacterium Aquaspirillum magnetotacticum is observed by the decay of the optical birefringence upon switching off an aligning magnetic field. The data yield a rotational diffusion constant Dr [unk] 0.13 s-1 and information about cell sizes that is consistent with optical microscopy data.  相似文献   

14.
The clinical course of a young woman following two separate suicide attempts using the herbicide paraquat is reported. The patient survived an intramuscular injection of paraquat almost asymptomatically, but later exhibited a typical fatal course with fulminant proliferative pulmonary fibrosis after an intravenous injection. Fibrosis and death occurred despite a therapeutic regimen based upon a known action of paraquat, the generation of superoxide (O2[unk]), using superoxide dismutase, α-tocopherol, and ascorbic acid in conjunction with forced diuresis and prednisone. While treatment failed explanations for the failure of therapy in this case and current therapeutic alternatives are discussed so that they may be considered when future cases are encountered.  相似文献   

15.
Specific immune precipitates dissolve in concentrated solutions of alkali-metal halides, and of alkaline-earth-metal halides and thiocyanates. The quantity of protein dissolved depends on the nature of the antigen-antibody system, on the proportion of the antigen in the precipitate, and on the avidity of the antibody. The extent of solubilization is a function of the temperature, of the volume of solution used and of the concentration of the ions in the solution, and also depends on the nature of these ions. The dissolving power of bivalent cations is greater than that of monovalent ones, and is as follows: Mg(2+)[unk]Ba(2+)[unk]Ca(2+)[unk]Sr(2+). Antigen-antibody complexes and free antibodies, but no free antigen, are detected in supernatants of specific precipitates dissolved in solutions of electrolytes of low ionic strength. Antigen-antibody complexes, free antibodies and also free antigen are detected in supernatants of specific precipitates dissolved in solutions of electrolytes of high ionic strength. Comparable results are obtained when the electrolyte solutions are studied for their effect on the bonds formed between an antibody and its corresponding immunosorbent. Moreover, in the latter case, 50% of the fixed antibodies could be recovered by elution with distilled water.  相似文献   

16.
1H and 13C nuclear-magnetic-resonance spectroscopy and functional-group analysis were used to determine the molecular structure of an isolated metabolite (IIb) of trimethyl-lysine as 3-hydroxy-N6-trimethyl-lysine, an important intermediate in the conversion of trimethyl-lysine into trimethylammoniobutyrate and carnitine [Hoppel, Cox & Novak (1980) Biochem. J. 188, 509–519]. Functional-group analysis revealed the presence of a primary amine and reaction of metabolite (IIb) with periodate yielded 4-N-trimethylammoniobutyrate as a product, showing 2,3-substitution on the molecule and suggesting that the 3-substitution on the molecule may be an alcohol ([unk]CH–OH), amine ([unk]CH[unk]–NH2) or carbonyl ([unk]C=O) functional group. 1H integration ratios, 1H and 13C chemical-shift data and 1H and 13C signal multiplicities from the sample (IIb) were used to complete the identification of metabolite (IIb) as 3-hydroxy-N6-trimethyl-lysine. For example, the proton multiplet at δ 4.2p.p.m. and doublet at δ 4.1p.p.m., positions representative of amine or alcohol substitution on methylene carbon atoms, integration ratios of 1:1:2:9:4 and a positive ninhydrin test suggest 3-hydroxy-N6-trimethyl-lysine as the molecular structure for metabolite (IIb). 13C chemical-shift data obtained from the sample (IIb) and compared with several model compounds (trimethylammoniohexanoate, trimethyl-lysine and 3-hydroxylysine) resulted in generation of the spectrum of the metabolite and allowed independent identification of metabolite (IIb) as 3-hydroxy-N6-trimethyl-lysine. The 1H spectrum of erythro- and threo-3-hydroxylysine are presented for comparison, and the 1H and 13C n.m.r. spectra of the erythro-isomer support this analysis.  相似文献   

17.
1. The reaction pathway for the carboxylation of pyruvate, catalysed by pig liver pyruvate carboxylase, was studied in the presence of saturating concentrations of K(+) and acetyl-CoA. 2. Free Mg(2+) binds to the enzyme in an equilibrium fashion and remains bound during all further catalytic cycles. MgATP(2-) binds next, followed by HCO(3) (-) and then pyruvate. Oxaloacetate is released before the random release, at equilibrium, of P(i) and MgADP(-). 3. This reaction pathway is compared with the double displacement (Ping Pong) mechanisms that have previously been described for pyruvate carboxylases from other sources. The reaction pathway proposed for the pig liver enzyme is superior in that it shows no kinetic inconsistencies and satisfactorily explains the low rate of the ATP[unk][(32)P]P(i) equilibrium exchange reaction. 4. Values are presented for the stability constants of the magnesium complexes of ATP, ADP, acetyl-CoA, P(i), pyruvate and oxaloacetate.  相似文献   

18.
The specifically 13C-labeled (90% 13C-enriched) peptide hormone derivatives [1-hem[2-13C]cystine]oxytocin, [1-hemi[1-13C]cystine]oxytocin, and [2-[-2-13C]tyrosine[-oxytocin and the analogue [3-[2-13C]leucine]oxytocin were prepared by total synthesis and used to study the interactions of the neurohypophyseal hormones with the bovine neurophysins as a function of pH and temperature. Under all conditions, whether high or low pH, the chemical shifts of the labeled carbon atoms of the bound hormones are the same, but they are shifted significantly from their positions in the free hormone. These results indicate that interactions of the side chain and disulfide moieties of the hormone with the neurophysins do not change as a function of pH. At neutral pH and 20--35 degrees C, the labeled atoms of the hormone are in slow exchange (1--5 s-1) with the neurophysins for the above hormone derivatives, but at low pH they are in intermediate or fast exchange depending upon the pH and temperature. At low pH, the dissociation rate constant (koff) is about 100-fold greater than the value at neutral pH, and this increase appears to be due exclusively to the breaking of the salt bridge involving the N-terminal amino group of oxytocin and a side-chain carboxyl group of neurophysin. Since the dissociation constant (Kd) also increases by about 100-fold in going from neutral to low pH, the association rate constant is deduced to be the same at neutral and low pH. In contrast to the low pH results, an increase in pH (from 6.6 to 10.5) leads to a continual decrease in the binding constant but to no apparent change in the dissociation rate constant. The bound hormone is always in slow exchange at high pH, even when the binding constant has been reduced by 2 or 3 orders of magnitude. At high pH, the decrease in binding affinity is due solely to the deprotonation of the alpha-amino group of the free hormone. Thus, at high pH the apparent association rate constant decreases, while the dissociation rate constant remains unchanged.  相似文献   

19.
Previous investigations on the persistence length of DNA in solution have revealed large discrepancies between hydrodynamic results and those from light-scattering techniques which have potentially a greater resolving power. The information obtained from experiments on a small circular DNA molecule has resolved these discrepancies. The non-superhelical circular double-stranded DNA molecule from bacteriophage [unk]X174-infected cells is small enough to permit accurate light-scattering extrapolations, and its solutions have negligible anisotropy. The persistence length obtained from experimental investigations on this molecule is comparable with that obtained by hydrodynamic techniques, even with variation of the excluded-volume factor.  相似文献   

20.
Pulsed field gradient NMR was used to measure the hydrodynamic behavior of unfolded variants of bovine pancreatic trypsin inhibitor (BPTI). The unfolded BPTI species studied were [R]Abu, at pH 4.5 and pH 2.5, and unfolded [14-38]Abu, at pH 2.5. These were prepared by chemical synthesis. [R]Abu is a model for reduced BPTI; all cysteine residues are replaced by alpha-amino-n-butyric acid (Abu). [14-38]Abu retains cysteines 14 and 38, which form a disulfide bond, while the other cysteine residues are replaced by Abu. In the PFG experiments, the diffusion coefficient is measured as a function of protein concentration, and the value of D degree -the diffusion coefficient extrapolated to infinite dilution-is determined. From D degree, a value of the hydrodynamic radius. Rh, is computed from the Stokes-Einstein relationship. At pH 4.5, [R]Abu has an Rh value significantly less than the value calculated for a random coil, while at pH 2.5 the experimental Rh value is the same as for a random coil. In view of the changes in NMR detected structure of [R]Abu at pH 4.5 versus pH 2.5 (Pan H, Barbar E, Barany G, Woodward C. 1995. Extensive non-random structure in reduced and unfolded bovine pancreatic trypsin inhibitor. Biochemistry 34:13974-13981), the collapse of reduced BPTI at pH 4.5 may be associated with the formation of non-native hydrophobic clusters of pairs of side chains one to three amino acids apart in sequence. The diffusion constant of [14-38]Abu was also measured at pH 4.5, where the protein is partially folded. An increase in hydrodynamic radius of partially folded [14-38]Abu, relative to native BPTI, is similar to the increase in radius of gyration measured for other proteins under "molten globule" conditions.  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