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 46 毫秒
1.
The reaction of perrhenate with 2-hydrazinopyrimidine in MeOH–HCl yields [ReCl31-NNC4H3N2H)(η2-HNNC4H3N2)] (1). The analogous reaction with Na2MoO4 yields [MoCl31-NNC4H3N2H)(η2-HNNHC4H3N2)] (1a). The reaction of 1 with pyrimidine-2-thiol and triethylamine produces [Re(η1-C4H3N2S)(η2-C4H3N2S)(η1-NNC4H3N2)(η2-HNNC4H3N2)] (2), while reaction of 1 with the Schiff base HSC6H4N=C(H)C6H4OH provides [Re(η3-SC6H4N=C(H)C6H4O)(η1-NNC4H3N2)(η2-HNNC4H3N2)]·0.6CH2Cl2 (3·0.6CH2Cl2). The analogous hydrazinopyridine complex of the Schiff base, [Re(η3-SC6H4N=C(H)C6H4O)(η1-NNC5H4N)(η2-HNNC5H4N)] (4), was also synthesized by reacting [ReCl31-NNC5H4NH)(η2-HNNC5H4N)] with HSC6H4N=C(H)C6H4OH. The crystal structures of 1–4 have been determined.  相似文献   

2.
Effect of iron concentration on hydrogen fermentation   总被引:11,自引:0,他引:11  
The effect of the iron concentration in the external environment on hydrogen production was studied using sucrose solution and the mixed microorganisms from a soybean-meal silo. The iron concentration ranged from 0 to 4000 mgFeCl2 l−1. The temperature was maintained at 37°C. The maximum specific hydrogen production rate was found to be 24.0 mlg−1 VSSh−1 at 4000 mgFeCl2 l−1. The specific production rate of butyrate increased with increasing iron concentration from 0 to 20 mgFeCl2 l−1, and decreased with increasing iron concentration from 20 to 4000 mgFeCl2 l−1. The maximum specific production rates of ethanol (682 mgg−1 VSSh−1) and butanol (47.0 mgg−1 VSSh−1) were obtained at iron concentrations of 5 and 3 mgFeCl2 l−1, respectively. The maximum hydrogen production yield of 131.9 mlg−1 sucrose was obtained at the iron concentration of 800 mgFeCl2 l−1. The maximum yields of acetate (389.3 mgg−1 sucrose), propionate (37.8 mgg−1 sucrose), and butyrate (196.5 mg g−1 sucros) were obtained at iron concentrations of 3, 200 and 200 mgFeCl2 l−1, respectively. The sucrose degradation efficiencies were close to 1.0 when iron concentrations were between 200 and 800 mgFeCl2 l−1. The maximum biomass production yield was 0.283 gVSSg−1 sucrose at an iron concentration of 3000 mgFeCl2 l−1.  相似文献   

3.
对缺刻缘绿藻(Parietochloris incisa(Reisigl) S.Watanabe)在不同光强和氮源及其浓度条件下的生长状况及油脂和花生四烯酸(AA)的积累规律进行了研究。结果显示,缺刻缘绿藻在3种氮源条件下均能较好地生长。在高氮浓度条件下,增大光强能显著提高缺刻缘绿藻的生物量并促进油脂和AA的积累。缺刻缘绿藻在300 μmol·m-2·s-1光强、8.8 mmol·L-1NaNO3条件下生物量达到最大(4.17 g·L-1)。油脂含量在100 μmol·m-2·s-1光强、1.0 mmol·L-1氮浓度下达到最高,分别为41.17%(NaNO3)、42.04%(NH4HCO3)和39.96%(CO(NH22)。AA绝对含量在300 μmol·m-2·s-1光强、2.9 mmol·L-1 NaNO3条件下达到最高,占细胞干重的16.44%。油脂和AA产率,在300 μmol·m-2·s-1光强、以NaNO3为氮源的条件下达到最大,分别为134.6 mg·L-1·d-1(1.0 mmol·L-1)和35.85 mg·L-1·d-1(2.9 mmol·L-1)。综合考虑成本等因素,选择NH4HCO3(5.9 mmol·L-1)和CO(NH22(2.9 mmol·L-1)为氮源、在300 μmol·m-2·s-1高光强下培养缺刻缘绿藻进行AA的生产为最优方案。  相似文献   

4.
为了探讨添加小麦秸秆和磷素对低磷土壤微生物数量和群落结构的影响,设置2个梯度的小麦秸秆添加量(N0和N1分别为0和2.08 g·kg-1)和4个施磷水平(P0、P1、P2和P3分别为0、100、200和400 mg·kg-1)组合处理,采用磷脂脂肪酸(PLFA)法测定土壤微生物生物量.结果表明: 添加秸秆配合施入磷素对微生物总生物量、细菌生物量、真菌生物量和真菌/细菌(F/B)比值具有显著的促进作用,微生物总生物量、细菌生物量、真菌生物量和F/B均为N1P1>N1P0>N1P2>N1P3>N0P1>N0P2>N0P3.在相同磷素水平下,添加秸秆处理的各指标均显著高于未添加秸秆处理;在添加相同秸秆量条件下,施磷处理的各指标随磷素施入量先增加后降低,以P1水平组合最优,其次是P0,最后是P2和P3.  相似文献   

5.
以野生红花白芨成熟蒴果为外植体,建立了白芨植株的再生体系。结果表明:白芨种子萌发最佳培养基为1/2 MS+6-BA 1.0 mg·L-1+NAA 0.1 mg·L-1+AC 0.5 g·L-1;原球茎增殖最适培养基为MS+KT 1.0 mg·L-1+NAA 0.1 mg·L-1+AC 0.5 g·L-1;原球茎分化培养基为MS+6-BA 0.5 mg·L-1+KT 1.0 mg·L-1+NAA 0.1 mg·L-1+AC 0.5 g·L-1;不定芽增殖最佳培养基为MS+6-BA 0.5 mg·L-1+KT 1.0 mg·L-1+NAA 0.1 mg·L-1+AC 0.5 g·L-1 +土豆汁150 g·L-1,30 d后的芽增殖率达343%;最适生根培养基为1/2MS+IBA 0.5 mg·L-1+AC 0.5 g·L-1+土豆汁100 g·L-1,生根率达100%;试管苗移栽的最佳方式是以水苔+树皮+椰糠+锯木粉(2∶2∶1∶1,V/V/V/V)为混合基质进行多株移栽,成活率可达99%。  相似文献   

6.
The kinetics and equilibria of complex formation by Ga(III) with NCS in aqueous solution have been measured over a range of acidities and temperatures, the contributing paths to the reaction resolved, and their rate constants and activation parameters determined. The hydrolysis equilibria required to carry out this resolution of kinetic behaviour have also been measured.

Unlike the other reported complexation reactions of Ga(III) in aqueous solution, the separate reaction pathways can be assigned with no ambiguity. At 25 °C and ionic strength 0.5 M, the observed forward rate constant for the complex formation is described by {k1 + k2K1h/[H+] + k3K1hK2h/[H+]2} M−1 s−1. For these conditions, the first and second successive hydrolysis constants of Ga(H2O)63+ are given by pK1h = 3.69 ± 0.01 and pK2h = 3.74 ± 0.04. The rate constants corresponding to the reactions of the species Ga(H2O)63+, Ga(H2O)5(OH)2+ and Ga(H2O)4(OH)2+ with NCS are k1 = 57 ± 4 M−1 −1, k2 = (1.08 ± 0.01) × 105 M−1 s−1 and k3 = 3 × 106 M−1 s−1 respectively. The complexation equilibrium quotient [GaNCS2+]/([Ga3+][NCS]) has been independently determined by spectrophotometric titration to be 20.8 ± 0.3 M−1 at 25 °C and ionic strength 0.5 M.

These kinetic results lead to an interpretation of the data, and a reinterpretation of other data for aquo-Ga(III) complex formation kinetics from the literature which support the assignment of a dissociative interchange mechanism for these reactions rather than the associative activation mode sometimes proposed.  相似文献   


7.
P.Muir Wood 《BBA》1974,357(3):370-379
The rate of electron transfer between reduced cytochrome ƒ and plastocyanin (both purified from parsley) has been measured as k = 3.6 · 107 M−1 · s−1, at 298 °K and pH 7.0, with activation parameters ΔH = 44 kJ · mole−1 and ΔS = +46 J · mole−1 · °K−1. Replacement of cytochrome ƒ with red algal cytochrome c-553, Pseudomonas cytochrome c-551 and mammalian cytochrome c gave rates at least 30 times slower: k = 5 · 105, 7.5 · 105 and 1.0 · 106 M−1 · s−1, respectively.

Similar measurements made with azurin instead of plastocyanin gave k = 6 · 106 and approx. 2 · 107 M−1 · s−1 for reaction of reduced azurin with cytochrome ƒ and algal cytochrome respectively.

Rate constants of 115 and 80 M−1 · s−1 were found for reduction of plastocyanin by ascorbate and hydroquinone at 298 °K and pH 7.0. The rate constants for the oxidation of plastocyanin, cytochrome ƒ, Pseudomonas cytochrome c-551 and red algal cytochrome c-553 by ferricyanide were found to be between 3 · 104 and 8 · 104 M−1 · s−1.

The results are discussed in relation to photosynthetic electron transport.  相似文献   


8.
O2 generation in mitochondrial electron transport systems, especially the NADPH-coenzyme Q10 oxidoreductase system, was examined using a model system, NADPH-coenzyme Q1-NADPH-dependent cytochrome P-450 reductase. One electron reduction of coenzyme Q1 produces coenzyme Q1 and O2 during enzyme-catalyzed reduction and O2 + coenzyme Q1 are in equilibrium with O2 + coenzyme Q1 in the presence of enough O2. The coenzyme Q1 produced can be completely eliminated by superoxide dismutase, identical to bound coenzyme Q10 radical produced in a succinate/fumarate couple-KCN-submitochondrial system in the presence of O2. Superoxide dismutase promotes electron transfer from reduced enzyme to coenzyme Q1 by the rapid dismutation of O2 generated, thereby preventing the reduction of coenzyme Q1 by O2. The enzymatic reduction of coenzyme Q1 to coenzyme Q1H2 via coenzyme Q1 is smoothly achieved under anaerobic conditions. The rate of coenzyme Q1H2 autoxidation is extremely slow, i.e., second-order constant for [O2][coenzyme Q1H2] = 1.5 M−1 · s−1 at 258 μM O2, pH 7.5 and 25°C.  相似文献   

9.
1. Rate constants for reduction of paraquat ion (1,1′-dimethyl-4,4′-bipyridy-lium, PQ2+) to paraquat radical (PQ+·) by eaq and CO2· have been measured by pulse radiolysis. Reduction by eaq is diffusion controlled (k = 8.4·1010 M−1·s−1) and reduction by CO2· is also very fast k = 1.5·1010 M−1·s−1).

2. The reaction of paraquat radical with oxygen has been analysed to give rate constants of 7.7·108 M−1·s−1 and 6.5·108 M−1·s−1 for the reactions of paraquat radical with O2 and O2·, respectively. The similarity in these rate constants is in marked contrast to the difference in redox potentials of O2 and O2· (− 0.59 V and + 1.12 V, respectively).

3. These rate constants, together with that for the self-reaction of O2·, have been used to calculate the steady-state concentration of O2· under conditions thought to apply at the site of reduction of paraquat in the plant cell. On the basis of these calculations the decay of O2· appears to be governed almost entirely by its self-reaction, and the concentration 5 μm away from the thylakoid is still 90% of that at the thylakoid itself. Thus, O2· persists long enough to diffuse as far as the chloroplast envelope and tonoplast, which are the first structures to be damaged by paraquat treatment. O2· is therefore sufficiently long-lived to be a candidate for the phytotoxic product formed by paraquat in plants.  相似文献   


10.
The free radical scavenging properties of retinyl ascorbate (RA-AsA) were determined by monitoring the decomposition of 2,2-diphenyl-1-picrylhydrazyl (DPPH) as a function of time and in comparison with ascorbic acid (AsA), ascorbic acid palmitate (AsA-Pal), retinoic acid (RA), retinol (ROL) and retinol palmitate (Rol-Pal). The rate constant of RA-AsA (mean3±SD) was 4.9±0.3 M-1 s-1, and indicated greater potency as an antioxidant compared to the rest of the test compounds (AsA 3.4±0.4 M-1 s-1, AsA-Pal, 2.9±0.2 M-1 s-1, RA 1.4±0.3 M-1 s-1, ROL 1.3±0.1 M-1 s-1, Rol-Pal exhibited insignificant activity). The decomposition rate constant of DPPH, 5±0.6 × 10-8 M-1 s-1, in ethanol and BHA, 154±3 M-1 s-1 were both used as control. The compound RA-2-carboxy-2-hydroxy-ethanoate was isolated by prep-TLC and was identified, by 13C and 1HNMR spectroscopy, as the major by-product from the reaction of RA-AsA with DPPH, which was also found to be potent antioxidant, 2.1±0.2 M-1 s-1. This suggests that oxidation of AsA moiety did not lead to the production of erythrulose species, which could cause deleterious modifications of cellular proteins.  相似文献   

11.
The kinetics of several processes involving the potential antioxidant role of urate in physiological systems have been investigated by pulse radiolysis. While the monoanionic urate radical, ·UH-, can be produced directly by oxidation with ·Br-2 or ·OH, it can also be generated by oxidation with the neutral tryptophan radical, ·Trp, with a rate constant of 2 × 107 M-1s-1. This radical, ·UH-, reacts with ·O-2 with a rate constant of 8 × 108 M-1s-1. Also, ·UH- is reduced by flavonoids, quercetin and rutin in CTAB micelles at rate constants of 6 × 106 M-1s-1 and 1 × 106 M-1s-1, respectively. These results can be of value by providing reference data useful in further investigation of the antioxidant character of urate in more complex biological systems.  相似文献   

12.
Addition of small amounts of Fe2+, Zn2+, Cu2+ and thiamine-HCl to the culture medium was required for promoting the galacto-oligosaccharide (Gal-OS)-producing activity of Sterigmatomyces elviae CBS8119, when the concentration of yeast extract in the medium was lowered to 0·1 g l−1. Galacto-oligosaccharide production using a recycling cell culture was performed in a medium containing 360 mg ml−1 of lactose supplemented with optimal concentrations of Fe2+ (1·5 mg l−1 of FeSO4.7H2O), Zn2+ (15 mg l−1 of ZnSO4.7H2O), Cu2+ (0·5 mg l−1 of CuSO4.5H2O) and thiamine-HCl (1 mg l−1 ) . Galacto-oligosaccharide production was maintained at high levels during six cycles of production, with the amount of Gal-OS produced in each cycle being more than 216 mg ml−1 (weight yield of more than 60%).  相似文献   

13.
Reactions of [(PPh3)2Pt(η3-CH2CCPh)]OTf with each of PMe3, CO and Br result in the addition of these species to the metal and a change in hapticity of the η3-CH2CCPh to η1-CH2CCPh or η1-C(Ph)=C=CH2. Thus, PMe3 affords [(PMe3)3Pt(η1-C(Ph)=C=CH2)]+, CO gives both [trans-(PPh3)2Pt(CO)(η1-CH2CCPh)]+ and [trans-(PPh3)2Pt(CO)(η1-C(Ph)=C=CH2)]+, and LiBr yields cis-(PPh3)2PtBr(η1-CH2CCPh), which undergoes isomerization to trans-(PPh3)2PtBr(η1-CH2CCPh). Substitution reactions of cis- and trans-(PPh3)2PtBr(η1-CH2CCPh) each lead to tautomerization of η1-CH2CCPh to η1-C(Ph)=C=CH2, with trans-(PPh3)2PtBr(η1-CH2CCPh) affording [(PMe3)3Pt(η1-C(Ph)=C=CH2)]+ at ambient temperature and the slower reacting cis isomer giving [trans-(PPh3)(PMe3)2Pt(η1-C(Ph)=C=CH2)]+ at 54 °C . All new complexes were characterized by a combination of elemental analysis, FAB mas spectrometry and IR and NMR (1H, 13C{1H} and 31P{1H}) spectroscopy. The structure of [(PMe3)3Pt(η1-C(Ph)=C=CH2)]BPh4·0.5MeOH was determined by single-crystal X-ray diffraction analysis.  相似文献   

14.
Yeast cytochrome c peroxidase (CCP) efficiently catalyzes the reduction of H2O2 to H2O by ferrocytochrome c in vitro. The physiological function of CCP, a heme peroxidase that is targeted to the mitochondrial intermembrane space of Saccharomyces cerevisiae, is not known. CCP1-null-mutant cells in the W303-1B genetic background (ccp1Δ) grew as well as wild-type cells with glucose, ethanol, glycerol or lactate as carbon sources but with a shorter initial doubling time. Monitoring growth over 10 days demonstrated that CCP1 does not enhance mitochondrial function in unstressed cells. No role for CCP1 was apparent in cells exposed to heat stress under aerobic or anaerobic conditions. However, the detoxification function of CCP protected respiring mitochondria when cells were challenged with H2O2. Transformation of ccp1Δ with ccp1W191F, which encodes the CCPW191F mutant enzyme lacking CCP activity, significantly increased the sensitivity to H2O2 of exponential-phase fermenting cells. In contrast, stationary-phase (7-day) ccp1Δ-ccp1W191F exhibited wild-type tolerance to H2O2, which exceeded that of ccp1Δ. Challenge with H2O2 caused increased CCP, superoxide dismutase and catalase antioxidant enzyme activities (but not glutathione reductase activity) in exponentially growing cells and decreased antioxidant activities in stationary-phase cells. Although unstressed stationary-phase ccp1Δ exhibited the highest catalase and glutathione reductase activities, a greater loss of these antioxidant activities was observed on H2O2 exposure in ccp1Δ than in ccp1Δ-ccp1W191F and wild-type cells. The phenotypic differences reported here between the ccp1Δ and ccp1Δ-ccp1W191F strains lacking CCP activity provide strong evidence that CCP has separate antioxidant and signaling functions in yeast.  相似文献   

15.
Carbonylation of the anionic iridium(III) methyl complex, [MeIr(CO)2I3] (1) is an important step in the new iridium-based process for acetic acid manufacture. A model study of the migratory insertion reactions of 1 with P-donor ligands is reported. Complex 1 reacts with phosphites to give neutral acetyl complexes, [Ir(COMe)(CO)I2L2] (L = P(OPh)3 (2), P(OMe)3 (3)). Complex 2 has been isolated and fully characterised from the reaction of Ph4As[MeIr(CO)2I3] with AgBF4 and P(OPh)3; comparison of spectroscopic properties suggests an analogous formulation for 3. IR and 31P NMR spectroscopy indicate initial formation of unstable isomers of 2 which isomerise to the thermodynamic product with trans phosphite ligands. Kinetic measurements for the reactions of 1 with phosphites in CH2Cl2 show first order dependence on [1], only when the reactions are carried out in the presence of excess iodide. The rates exhibit a saturation dependence on [L] and are inhibited by iodide. The reactions are accelerated by addition of alcohols (e.g. 18× enhancement for L = P (OMe)3 in 1:3 MeOH-CH2Cl2). A reaction mechanism is proposed which involves substitution of an iodide ligand by phosphite, prior to migratory CO insertion. The observed rate constants fit well to a rate law derived from this mechanism. Analysis of the kinetic data shows that k1, the rate constant for iodide dissociation, is independent of L, but is increased by a factor of 18 on adding 25% MeOH to CH2Cl2. Activation parameters for the k1 step are ΔH = 71 (±3) kJ mol, ΔS = −81 (±9) J mol−1 K−1 in CH2Cl2 and ΔH = 60(±4) kJ mol−1, ΔS = −93(± 12) J mol−1 K−1 in 1:3 MeOH-CH2Cl2. Solvent assistance of the iodide dissociation step gives the observed rate enhancement in protic solvents. The mechanism is similar to that proposed for the carbonylation of 1.  相似文献   

16.
Yin J  Vogel U  Ma Y  Qi R  Wang H 《Mutation research》2008,641(1-2):12-18
To evaluate the joint effect of nine single nucleotide polymorphisms for three DNA repair genes in the region of chromosome 19q13.2-3 on susceptibility of lung cancer in a Chinese population, we conducted a hospital-based case–control study consisting of 247 lung cancer cases and 253 cancer-free controls matched on age, gender and ethnicity. Associations between the haplotypes and susceptibility of lung cancer were tested. The global test of haplotype association revealed a statistically significant difference in the haplotype distribution between cases and controls (global test: χ2 = 60.45, d.f. = 15, P = 2.11E−07). The two haplotypes were underrepresented among cases (Hap5 defined by ERCC1118AERCC2156CERCC2312GERCC2751AXRCC1194TXRCC1206AXRCC1280GXRCC1399GXRCC1632G and Hap12 defined by ERCC1118GERCC2156CERCC2312GERCC2751AXRCC1194CXRCC1206AXRCC1280GXRCC1399AXRCC1632G). Three of the haplotypes were overrepresented among cases (Hap3 defined by ERCC1118AERCC2156CERCC2312GERCC2751AXRCC1194CXRCC1206AXRCC1280GXRCC1399GXRCC1632G, Hap4 defined by ERCC1118AERCC2156CERCC2312GERCC2751AXRCC1194CXRCC1206GXRCC1280GXRCC1399GXRCC1632A, and Hap10 defined by ERCC1118GERCC2156AERCC2312GERCC2751AXRCC1194TXRCC1206AXRCC1280GXRCC1399GXRCC1632G). Haplotypes 3 and 10 (cases = 5.7%, controls = 1.0%, OR = 6.56, 95%CI = 1.83–23.54, P = 0.001; cases = 13.3%, controls = 5.6%, OR = 2.73, 95%CI = 1.51–4.94, P = 0.0006) were the most strongly associated with increased lung cancer risk. There was considerable linkage disequilibrium exists between SNPs both within genes and between genes in the region. The two blocks for solid spine of LD and six htSNPs were found. The haplotype analysis suggested that the biologically effective polymorphisms co-segregate with some of the haplotypes. This result supports the hypothesis that the sub-region is important for lung cancer susceptibility. Haplotype studies using larger study groups will be required to obtain conclusive results.  相似文献   

17.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

18.
白檀离体快繁技术   总被引:1,自引:0,他引:1  
以白檀(Symplocos paniculata)幼嫩茎段为实验材料, 通过对启动培养、增殖、生根培养及移栽的影响因子进行研究, 初步建立了白檀的组织培养体系。结果表明: 白檀外植体最适灭菌方案为0.1%升汞3分钟, 无菌苗获得率达81%; 最适初代启动培养基为1/2MS+30 g∙L-1蔗糖+8 g∙L-1琼脂, 出芽率达86.83%; 增殖最适培养基为1/2MS+1.0 mg∙L-1 6-BA+0.02 mg∙L-1 IBA+30 g∙L-1蔗糖+8 g∙L-1琼脂, 增殖系数达3.57; 最适生根培养基为WPM+0.5 mg∙L-1 IBA+0.5 mg∙L-1 NAA+20 g∙L-1蔗糖+2 g∙L-1 AC+8 g∙L-1琼脂, 生根率达93%; 炼苗后, 移入园土:草炭土=1:1 (v/v)的基质中, 成活率达83%。  相似文献   

19.
利用田间小区试验,设计生物炭用量为0(B0)、1 kg·m-2(B1)、2 kg·m-2(B2)3个水平,氮肥用量为0(N0)、40 g·m-2(N1)、60 g·m-2(N2)3个水平, 即B0N0、B0N1、B0N2、B1N0、B1N1、B1N2、B2N0、B2N1和B2N2共9个处理,研究了生物炭与氮肥配施对牡丹叶片的氮素积累、叶片氮素向籽粒转移、籽粒蛋白氮、氨基酸和脂肪酸含量,以及籽粒产量和品质的影响.结果表明: 生物炭与氮肥配施增加了牡丹不同发育时期叶片中蛋白氮和非蛋白氮含量,以及叶片氮素向籽粒的转移量和籽粒氮素积累量.与B0N0处理相比,B1N1处理叶片氮素转移量和籽粒氮素积累量分别增加27.6%和27.1%;B1N1和B2N1处理牡丹籽粒百粒重和籽粒产量分别增加13.6%和16.4%,其中籽粒产量在B1N1、B1N2、B2N1和B2N2处理间差异不显著;B2N1和B1N2处理牡丹籽粒蛋白氮和总氨基酸含量较高,分别增加29.3%和36.2%.生物炭与氮肥配施增加了牡丹籽粒中总脂肪酸和不饱和脂肪酸的含量,其中,B2N1处理总脂肪酸含量较高, 比B0N0处理增加了17.4%.生物炭与氮肥配施能够增加牡丹叶片氮素积累量和叶片氮素向籽粒的转移量,增加籽粒产量,提高牡丹籽粒蛋白氮、氨基酸和脂肪酸的含量,其中以生物炭1 kg·m-2与氮肥40 g·m-2配施效果较好.  相似文献   

20.
Experimental evidence is provided that selenomethionine oxide (MetSeO) is more readily reducible than its sulfur analogue, methionine sulfoxide (MetSO). Pulse radiolysis experiments reveal an efficient reaction of MetSeO with one-electron reductants, such as e-aq (k = 1.2 × 1010M-1s-1), CO·-2 (k = 5.9 × 108 M-1s-1) and (CH3)2) C·OH (k = 3.5 × 107M-1s-1), forming an intermediate selenium-nitrogen coupled zwitterionic radical with the positive charge at an intramolecularly formed Se N 2σ/1σ* three-electron bond, which is characterized by an optical absorption with λmax at 375 nm, and a half-life of about 70 μs. The same transient is generated upon HO· radical-induced one-electron oxidation of selenomethionine (MetSe). This radical thus constitutes the redox intermediate between the two oxidation states, MetSeO and MetSe. Time-resolved optical data further indicate sulfur-selenium interactions between the Se N transient and GSH. The Se N transient appears to play a key role in the reduction of selenomethionine oxide by glutathione.  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