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 15 毫秒
1.
Yields based on carbon are usually reported in prebiotic experiments, while energy yields (moles cal-1) are more useful in estimating the yields of products that would have been obtained from the primitive atmosphere of the earth. Energy yields for the synthesis of HCN and H2CO from a spark discharge were determined for various mixtures of CH4, CO, CO2, H2, H2O, N2 and NH3. The maximum yields of HCN and H2CO from CH4, CO, and CO2 as carbon sources are about 4 X 10(-8) moles cal-1.  相似文献   

2.
The molybdenum-dependent nitrogenase catalyzes the multi-electron reduction of protons and N(2) to yield H(2) and 2NH(3). It also catalyzes the reduction of a number of non-physiological doubly and triply bonded small molecules (e.g. C(2)H(2), N(2)O). Carbon monoxide (CO) is not reduced by the wild-type molybdenum nitrogenase but instead inhibits the reduction of all substrates catalyzed by nitrogenase except protons. Here, we report that when the nitrogenase MoFe protein α-Val(70) residue is substituted by alanine or glycine, the resulting variant proteins will catalyze the reduction and coupling of CO to form methane (CH(4)), ethane (C(2)H(6)), ethylene (C(2)H(4)), propene (C(3)H(6)), and propane (C(3)H(8)). The rates and ratios of hydrocarbon production from CO can be adjusted by changing the flux of electrons through nitrogenase, by substitution of other amino acids located near the FeMo-cofactor, or by changing the partial pressure of CO. Increasing the partial pressure of CO shifted the product ratio in favor of the longer chain alkanes and alkenes. The implications of these findings in understanding the nitrogenase mechanism and the relationship to Fischer-Tropsch production of hydrocarbons from CO are discussed.  相似文献   

3.
Catalytic properties of aluminosilicates may play a role in the synthesis of biological molecules from simple gaseous molecules commonly found in planetary atmospheres. Urea, amino acids and UV absorbing substances have been obtained by heating CO and NH3 with Linde molecular sieves saturated with Ca+2, NH4+ or Fe+3. The yields of amino acids produced have been determined by an amino acid analyzer. The quantity of urea produced largely depends on the nature of the saturating cation. Experiments using 14CO confirm that the amino acids are not due to contaminants adsorbed on the surface of the molecular sieves.  相似文献   

4.
Reaction of elemental copper and zinc powder mixtures with glycine (NH2.CH2COOH; HA) or aspartic acid (NH2CHCOOHCH2COOH; H2B) (in 1:1:2 ratio, respectively) in the presence of excess hydrogen peroxide (H2O2) at 50 degrees C, results in the formation of a new mixed metal peroxy carbonate compound corresponding to formula [Cu(Zn)2(O2(2-) (CO3)2(H2O)4], while the same reaction with elemental copper powder alone yields merely peroxy amino acid compounds having the formula [Cu(O2(2-)) (HA)2(H2O)] and [Cu(O2(2-)) (H2B) (H2O)2] for glycine and aspartic acid, respectively. These compounds have been characterized by elemental analysis, ESR, and electronic and IR spectra. It is interesting to note that both amino acids are converted to carbonate in the presence of zinc alone. A method analogous to that described above, for the reaction of elemental copper, zinc powder mixtures with succinic acid [(CH2COOH)2] or acetic acid (CH3COOH) in excess H2O2, on the other hand, gave a product essentially comprising copper succinate or acetate, respectively. These observations suggest an interesting and perhaps important phenomenon by which only the simple amino acids such as glycine and aspartic acid are converted to carbonates while their corresponding carboxylic acids form only their respective salts.  相似文献   

5.
Hydrazinonicotinamide (HYNIC) forms stable coordination complexes with Tc-99m when reacted with Tc(V)oxo species such as Tc-mannitol or other Tc-polyhydric complexes. However, radio-HPLC of [Tc-For-MLFK-HYNIC] labeled via Tc-polyhydric ligands demonstrated multiple radiochemical species each with unique biodistribution patterns. This is likely due to the fact that Tc can bind to the hydrazino moiety, as well as polyhydric ligands, in a variety of coordination geometries. Tridentate ligands, such as bis(mercaptoethyl)methylamine (NS2), may constrain the possible coordination geometries and improve overall stability. To investigate this, we synthesized NS2, converted the [Tc-mannitol-For-MLFK-HYNIC] to the corresponding NS2-containing complex [Tc-NS2-For-MLFK-HYNIC], and compared its infection imaging and biodistribution properties with [Tc-mannitol-For-MLFK-HYNIC]. Conversion to the NS2 complex was confirmed by HPLC which showed a single unique hydrophobic species with retention time greater than the [Tc-mannitol-For-MLFK-HYNIC] complex. Imaging experiments with both preparations were performed in rabbits with E. coli infections in the left thigh. Tissue radioactivity measurements demonstrated that compared to Tc-mannitol-peptide, accumulation of Tc-NS2-peptide was lower in blood, heart, and normal muscle and higher in spleen, infected muscle, and pus (p < 0.01). These results indicate that the Tc-NS2-peptide complex is chemically more homogeneous and exhibits improved infection localization and biodistribution properties. In an effort to model the interactions of the metal-HYNIC core with NS2 and related ligand types, the reactions of [ReCl3(NNC5H4NH)(NHNC5H4N)] and [99TcCl3(NNC5H4NH)(NHNC5H4N)], effective structural analogues for the [M(NNC5H4NH(x))2] core, with NS2, C5H3N-2,6-(CH2SH)2, O(CH2CH2SH)2, and S(CH2CH2SH)2 were investigated and the compounds [M[CH3N(CH2CH2S)2](NNC5H4N)(NHNC5H4N] (M = 99Tc (5a), Re (5b)), [Re[C5H3N-2,6-(CH2S)2](NNC5H4N)(NHNC5H4N)].CH2Cl2.0.5MeOH (7), [Re[SCH2CH2)2O] (NNC5H4N)(NHNC5H4N)] (8), and [Re[(SCH2CH2)2S](NNC5H4NH)(NHNC5H4N)]Cl (9) were isolated. Similarly, the reaction of [ReCl3(NNC5H4NH)(NHNC5H4N)] with the bidentate ligands pyridine-2-methanethiol and 3-(trimethlysilyl)pyridine-2-thiol led to the isolation of [ReCl(C5H4N-2-CH2S) (NNC5H4N)(NHNC5H4N)] (10) and [Re(2-SC5H3N-3-SiMe3)2 (NNC5H4N)(NHNC5H4N)] (11), respectively, while reaction with N-methylimidazole-2-thiol yielded the binuclear complex [Re(OH)Cl(SC3H2N2CH3)2(NNC5H4N)2 (NHNC5H4N)2] (12). The analogous metal-(HYNIC-OH) precursor, [ReCl3[NNC5H3NH(CO2R)] [NHNC5H3N(CO2R)]] (R = H, 13a; R = CH3, 13b) has been prepared and coupled to lysine to provide [RCl3[NNC5H3NH(CONHCH2CH2CH2CH2CH(NH2)CO2H)] [NHNC5H3NH(CONHCH2CH2CH2CH2CH(NH2)CO2H)]].2HCl (14.2HCl), while the reaction of the methyl ester 13b with 2-mercaptopyridine yields [Re(2-SC5H4N)2[NNC5H3N(CO2Me)][NHNC5H3N(CO2Me)]] (15). While the chemical studies confirm the robustness of the M-HYNIC core (M = Tc, Re) and its persistence in ligand substitution reactions at adjacent coordination sites of the metal, the isolation of oligomeric structures and the insolubility of the peptide conjugates of 13, 14, and 15 underscore the difficulty of characterizing these materials on the macroscopic scale, an observation relevant to the persistent concerns with reagent purity and identity on the tracer level.  相似文献   

6.
(1) Cyanamide (N identical to C-NH2) has been shown to be a substrate for purified Mo-nitrogenases of Klebsiella pneumoniae and Azotobacter chroococcum, with apparent Km values near 0.8 mM. (2) Reduction products were CH4, CH3NH2 and NH3 formed by pathways requiring 6 or 8 electrons: N identical to CNH2 + 6e + 6H+----CH3NH2 + NH3; N identical to CNH2 + 8e + 8H+----CH4 + 2NH3 (3) Acetylene reduction and hydrogen evolution were inhibited more than 75% by cyanamide (10 mM). Cyanamide also inhibited total electron flux at nitrogenase protein component ratios (Fe/MoFe) near 10. (4) Cyanamide was also a substrate for the recently isolated Va-nitrogenase of A. chroococcum, but with an apparent Km of 2.6 mM showed weaker binding and an 8-fold lower Vmax than did either Mo-nitrogenase. (5) The component ratios of nitrogenase proteins favouring CH4 formation was 3.5 Fe/MoFe protein and 1 Fe/VaFe protein.  相似文献   

7.
Cell suspensions of methanogenic bacteria (Methanosarcina barkeri, Methanospirillum hungatei, Methano-brevibacter arboriphilus, and Methanobacterium thermoautotrophicum) were found to form CO from CO2 and H2 according to the reaction: CO2 + H2----CO + H2O; delta G0 = +20 kJ/mol. Up to 15,000 ppm CO in the gas phase were reached which is significantly higher than the equilibrium concentration calculated from delta G0 (95 ppm under the experimental conditions). This indicated that CO2 reduction with H2 to CO is energy-driven and indeed the cells only generated CO when forming CH4. The coupling of the two reactions was studied in more detail with acetate-grown cells of M. barkeri using methanogenic substrates. The effects of the protonophore tetrachlorosalicylanilide (TCS) and of the proton-translocating ATPase inhibitor N,N'-dicyclohexylcarbodiimide (cHxN)2C were determined. TCS completely inhibited CO formation from CO2 and H2 without affecting methanogenesis from CH3OH and H2. In the presence of the protonophore the proton motive force delta p and the intracellular ATP concentration were very low. (cHxN)2C, which partially inhibited methanogenesis from CH3OH and H2, had no effect on CO2 reduction to CO. In the presence of (cHxN)2C delta p was high and the intracellular ATP content was low. These findings suggest that the endergonic formation of CO from CO2 and H2 is coupled to the exergonic formation of CH4 from CH3OH and H2 via the proton motive force and not via ATP. CO formation was not stimulated by the addition of sodium ions.  相似文献   

8.
Ammonia oxidizers (family Nitrobacteraceae) and methanotrophs (family Methylococcaceae) oxidize CO and CH4 to CO2 and NH4+ to NO2-. However, the relative contributions of the two groups of organisms to the metabolism of CO, CH4, and NH4+ in various environments are not known. In the ammonia oxidizers, ammonia monooxygenase, the enzyme responsible for the conversion of NH4+ to NH2OH, also catalyzes the oxidation of CH4 to CH3OH. Ammonia monooxygenase also mediates the transformation of CH3OH to CO2 and cell carbon, but the pathway by which this is done is not known. At least one species of ammonia oxidizer, Nitrosococcus oceanus, exhibits a Km for CH4 oxidation similar to that of methanotrophs. However, the highest rate of CH4 oxidation recorded in an ammonia oxidizer is still five times lower than rates in methanotrophs, and ammonia oxidizers are apparently unable to grow on CH4. Methanotrophs oxidize NH4+ to NH2OH via methane monooxygenase and NH4+ to NH2OH via methane monooxygenase and NH2OH to NO2- via an NH2OH oxidase which may resemble the enzyme found in ammonia oxidizers. Maximum rates of NH4+ oxidation are considerably lower than in ammonia oxidizers, and the affinity for NH4+ is generally lower than in ammonia oxidizers. NH4+ does not apparently support growth in methanotrophs. Both ammonia monooxygenase and methane monooxygenase oxidize CO to CO2, but CO cannot support growth in either ammonia oxidizers or methanotrophs. These organisms have affinities for CO which are comparable to those for their growth substrates and often higher than those in carboxydobacteria. The methane monooxygenases of methanotrophs exist in two forms: a soluble form and a particulate form. The soluble form is well characterized and appears unrelated to the particulate. Ammonia monooxygenase and the particulate methane monooxygenase share a number of similarities. Both enzymes contain copper and are membrane bound. They oxidize a variety of inorganic and organic compounds, and their inhibitor profiles are similar. Inhibitors thought to be specific to ammonia oxidizers have been used in environmental studies of nitrification. However, almost all of the numerous compounds found to inhibit ammonia oxidizers also inhibit methanotrophs, and most of the inhibitors act upon the monooxygenases. Many probably exert their effect by chelating copper, which is essential to the proper functioning of some monooxygenases. The lack of inhibitors specific for one or the other of the two groups of bacteria hampers the determination of their relative roles in nature.  相似文献   

9.
We present a detailed analysis of the variety, quantity and distribution of the amino acids detected in organic residues after acid hydrolysis. Such organic residues are produced in the laboratory after the vacuum ultraviolet (VUV) irradiation of several astrophysically relevant ice mixtures containing H(2)O, CO, CO(2), CH(3)OH, CH(4) and NH(3) at low temperature (10-80 K), and subsequent warm-up to room temperature. We explore five experimental parameters: the irradiation time, the temperature, the ice mixture composition, the photon dose per molecule and the substrate for the ice deposition. The amino acids were detected and identified by ex-situ liquid chromatography analysis of the organic residues formed after warming the photolysed ices up to room temperature. This study shows that in all experiments amino acids are formed. Their total quantities and distribution depend slightly on the experimental parameters explored in the present work, the important requirement to form such molecules being that the starting ice mixtures must contain the four elements C, H, O and N. We also discuss the effects of the chemical treatment needed to detect and identify the amino acids in the organic residues. Finally, these results are compared with meteoritic amino acid data from the carbonaceous chondrite Murchison, and the formation processes of such compounds under astrophysical conditions are discussed.  相似文献   

10.
Radiolabeling of biologically active molecules with the [(99m)Tc(CO)(3)](+) unit has been of primary interest in recent years. With this in mind, we herein report symmetric (L(1)) and asymmetric (L(2)-L(5)) pyrazolyl-containing chelators that have been evaluated in radiochemical reactions with the synthon [(99m)Tc(H(2)O)(3)(CO)(3)](+) (1a). These reactions yielded the radioactive building blocks [(99m)Tc(CO)(3)(k(3)-L)](+) (L = L(1)-L(5), 2a-6a), which were identified by RP-HPLC. The corresponding Re surrogates (2-6) allowed for macroscopic identification of the radiochemical conjugates. Complexes 2a-6a, with log P(o/w) values ranging from -2.35 to 0.87, were obtained in yields of > or =90% using ligand concentrations in the 10(-5-)10(-4) M range. Challenge studies with cysteine and histidine revealed high stability for all of these radioactive complexes, and biodistribution studies in mice indicated a fast rate of blood clearance and high rate of total radioactivity excretion, occurring primarily through the renal-urinary pathway. Based on the framework of the asymmetric chelators, the novel bifunctional ligands 3,5-Me(2)-pz(CH(2))(2)N((CH(2))(3)COOH)(CH(2))(2)NH(2) (L(6)) and pz(CH(2))(2)N((CH(2))(3)COOH)(CH(2))(2)NH(2) (L(7)) have been synthesized and their coordination chemistry toward (NEt(4))(2)[ReBr(3)(CO)(3)] (1) has been explored. The resulting complexes, fac-[Re(CO)(3)(k(3)-L)]Br (L(6)(7), L(7)(8)), contain tridentate ancillary ligands that are coordinated to the metal center through the pyrazolyl and amine nitrogen atoms, as observed for the other related building blocks. L(6) and L(7) were coupled to a glycylglycine ethyl ester dipeptide, and the resulting functionalized ligands were used to prepare the model complexes fac-[Re(CO)(3)(kappa(3)-3,5-Me(2)-pz(CH(2))(2)N(glygly)(CH(2))(2)NH(2))](+) (9/9a) and fac-[Re(CO)(3)(kappa(3)-pz(CH(2))(2)N(CH(2))(3)(glygly)(CH(2))(2)NH(2))](+) (10/10a) (M = Re, (99m)Tc). These small conjugates have been fully characterized and are reported herein. On the basis of the in vitro/in vivo behavior of the model complexes (2a-6a, 9a, 10a), we chose to evaluate the in vitro/in vivo biological behavior of a new tumor-seeking Bombesin pyrazolyl conjugate, [(L(6))-G-G-G-Q-W-A-V-G-H-L-M-NH(2)], that has been labeled with the [(99m)Tc(CO)(3)](+) metal fragment. Stability, in vitro cell binding assays, and pharmacokinetics studies in normal mice are reported herein.  相似文献   

11.
Free amino acid turnover in methanogens measured by 15N NMR spectroscopy   总被引:1,自引:0,他引:1  
Turnover of the nitrogen moiety from free amino acid pools in two thermophilic methanogens, Methanobacterium thermautotrophicum delta H and Methanococcus thermolithotrophicus SN1, has been monitored with 15N NMR spectroscopy. In cells growing exponentially on 15NH4Cl, glutamate was the major soluble 15N-labeled species in both organisms. When the Mb. thermoautotrophicum cells were harvested, washed, and resuspended into medium containing 14NH4Cl, the resonance for [15N]glutamate decreased with a half-life of 0.5 h. This is considerably faster than the turnover rate for the carbon side chain of glutamate (7 h) obtained when a 13CO2 pulse followed by a 12CO2 chase was incorporated into the 15N/14N-labeling experiment. Such behavior is consistent with recycling of the glutamate carbon skeleton via alpha-ketoglutarate after transamination reactions remove the 15N for biosynthesis of other amino acids, nucleic acids, etc. When the cells were in stationary phase, 15N turnover was considerably slower indicating that transaminase activity had also decreased. Mc. thermolithotrophicus has a much more fragile cell wall and easily lyses. To avoid cell loss in the 15N/14N experiment, 15NH+4 growth followed by 14NH4+ dilution was used. In this organism the glutamate-labeled nitrogen turns over quite rapidly (t1/2 approximately 9 min), at a rate comparable to that for the carbon skeleton (t1/2 approximately 10 min). Beta-Glutamate, the second major carbon and nitrogen pool in this organism, turns over its 15N label very slowly. Therefore, this beta-amino acid does not appear to serve as a nitrogen donor in Mc. thermolithotrophicus.  相似文献   

12.
Geometry optimization and energy calculations have been performed at the density functional B3LYP/LANL2DZ level on hydrogen sulfide (HS-), dihydrogensulfide (H2S), thiomethanolate (CH3S-), thiomethanol (CH3SH), thiophenolate (C6H5S-), methoxyde (CH3O-), methanol (CH3OH), formiate (HCOO-), acetate (CH3COO-), carbonate (CO3(2-)), hydrogen carbonate (HCO3-), iminomethane (NH=CH2), [ZnS], [ZnS2]2-, [Zn(HS)]+, [Zn(H2S)]2+, [Zn(HS)4]2-, [Zn(CH3S)]+, [Zn(CH3S)2], [Zn(CH3S)3]-, [Zn(CH3S)4]2-, [Zn(CH3SH)]2+, [Zn(CH3SCH3)]2+, [Zn(C6H5S)]+, [Zn(C6H5S)2], [Zn(C6H5S)3]-, [Zn(HS)(NH=CH2)2]+, [Zn(HS)2(NH=CH2)2], [Zn(HS)(H2O)]+, [Zn(HS)(HCOO)], [Zn(HS)2(HCOO)]-, [Zn(CH3O)]+, [Zn(CH3O)2], [Zn(CH3O)3]-, [Zn(CH3O)4]2, [Zn(CH3OH)]2+, [Zn(HCOO)]+, [Zn(CH3COO)]+, [Zn(CH3COO)2], [Zn(CH3COO)3]-, [Zn(CO3)], [Zn(HCO3)]+, and [Zn(HCO3)(Imz)]+ (Imz, 1,3-imidazole). The computed Zn-S bond distances are 2.174A for [ZnS], 2.274 for [Zn(HS)]+, 2.283 for [Zn(CH3S)]+, and 2.271 for [Zn(C6H5S)]+, showing that sulfide anion forms stronger bonds than substituted sulfides. The nature of the substituents on sulfur influences only slightly the Zn-S distance. The optimized tetra-coordinate [Zn(HS)2(NH=CH2)2] molecules has computed Zn-S and Zn-N bond distances of 2.392 and 2.154A which compare well with the experimental values at the solid state obtained via X-ray diffraction for a number of complex molecules. The computed Zn-O bond distances for chelating carboxylate derivatives like [Zn(HOCOO)]+ (1.998A), [Zn(HCOO)]+ (2.021), and [Zn(CH3COO)]+ (2.001) shows that the strength of the bond is not much influenced by the substituent on carboxylic carbon atom and that CH3- and HO- groups have very similar effects. The DFT analysis shows also that the carboxylate Ligand has a preference for the bidentate mode instead of the monodentate one, at least when the coordination number is small.  相似文献   

13.
13C, 15N, and 1H nuclear magnetic resonance measurements indicate that chloroform-soluble threonine-containing tripeptide derivatives, such as t-Boc-Thr-Gly-Gly-OBz, form three strong hydrogen bonds to the cytosine moiety of 2',3'-O-isopropylidene-5'-O-t-butyldimethylsilylcytidine. The C = O and NH of the central peptide residue plus the OH of the threonine side chain appear to form bonds to the N(4')H2, N(3), and C(2) = O, respectively, of the pyrimidine. An association constant calculated from the cytidine 15N(4') nuclear magnetic resonance response to added peptide is four times larger than the corresponding cytosine-guanine constant. It is suggested that cytosine-peptide bonding was part of the primitive genetic coding mechanism early in evolution and accounts for the origin of the cytosine-centered codons for the hydroxy amino acids, serine and threonine, in the present code.  相似文献   

14.
(+)-18-crown-6 tetracarboxylic acid (18C6H(4)) has been used as a chiral selector for various amines and amino acids. To further clarify the structural scaffold of 18C6H(4) for chiral separation, single crystal X-ray analysis of its glycine(+) (1), H3O+ (2), H5O2+ (3), NH4+ (4), and 2CH3NH3+ (5) complexes was performed and the guest-dependent conformation of 18C6H(4) was investigated. The crown ether ring of 18C6H4 in 3, 4, and 5 took a symmetrical C2 or C2-like conformation, whereas that in 1 and 2 took an asymmetric C1 conformation, which is commonly observed in complexes with various optically active amino acids. The overall survey of the present and related complexes suggests that the molecular conformation of 18C6H4 is freely changeable within an allowable range, depending on the molecular shape and interaction mode with the cationic guest. On the basis of the present results, we propose the allowable conformational variation of 18C6H4 and a possible transition pathway from its primary conformation to the conformation suitable for chiral separation of racemic amines and amino acids.  相似文献   

15.
The promoting activity of polyamine analogs (IV approximately XV) on staphylococcal nuclease with DNA as the substrate was compared with that of natural polyamines (I APPROXIMATELY III): I. NH2(CH2)3NH(CH2)4NH(CH2)3NH2(spermine); II. NH2(CH2)3NH(CH2)3NH(CH2)3NH2(thermine); III. NH2(CH2)4NH2 (putrescine); IV. CN(CH2)2NH(CH2)4NH(CH2)2CN; V. HOOC(CH2)2NH(CH2)4NH(CH2)2COOH; VI. C2H5OOC(CH2)2NH(CH2)4NH(CH2)2COOC2H5; VII. HO(CH2)3NH(CH2)4HH(CH2)3OH; VIII. CH3COHH(CH2)3NH(CH2)4NH(CH2)3NHCOCH3; IX. C2H5NH(CH2)3NH(CH2)4NH(CH2)3NHC2H5; X. NH2(CH2)3S(CH2)4S(CH2)3NH2; XI. NH2(CH2)3NH(CH2)2O(CH2)2NH(CH2)3NH2; XII. NH2(CH2)3NCH3(CH2)4HCH3(CH2)3NH2; XIII. CN(CH2)2NCH3(CH2)4NCH3(CH2)2CN; XIV. (CH3)2N(CH2)3NCH3(CH2)4NCH3(CH2)3N(CH3)2; XV. NH2(CH2)2O(CH2)2NH2 Replacement of the terminal groups by CN, COOH, COOEt, NHAc, NHEt, or N(CH3)2 remarkably decreased the activity. The compound VII with terminal hydroxyl groups had a lower promoting activity at low concentrations, but revealed higher activity at higher concentrations and, in contrast to spermine, no inhibition at all even at very high concentrations. Replacement of both internal amino groups by sulfur or NCH3 decreased the activity. The introduction of an ether bond into the internal methylene groups (compound XI) highly decreased the activity. Based upon these findings the possible relationship between structure and activity is discussed.  相似文献   

16.
The reaction of [VO(CH3COO)2(phen)] (phen = 1,10-phenanthroline) with the sulfhydryl-containing pseudopeptides (scp), N-(2-mercaptopropionyl)glycine (H3mpg), N-(2-mercaptopropionyl)cysteine (H4m2pc), N-(3-mercaptopropionyl)cysteine (H4m3pc) and the dipeptides glycylglycine (H2glygly) and glycyl-L-alanine (H2glyala), in the presence of triethylamine, results in the formation of the compounds Et3NH[VO(mpg)(phen)] (1), (Et3NH)2[VO(m2pc)] (4), [(Et3NH)2[VO(m3pc) (5), [VO(glygly)(phen)] x 2CH3OH (2 x 2CH3OH) and [VO(glyala)(phen)] x CH3OH (3 x CH3OH). Evidence for the molecular connectivity in 2 x CH3OH was established by X-ray crystallography, showing the vanadium(IV) atom ligated to a tridentate glygly2- ligand at the N(amine), N(peptide) and O(carboxylato) atoms. Combination of the correlation plot of the EPR parameters gz versus Az, together with the additivity relationship supported the prediction of the equatorial donor atom sets of the V(IV)O2+ center at various pH values for the V(IV)O2+-glutathione system considered in this study. Model NMR studies (interaction of vanadium(V) with the scp H3mpg) showed that there is a possibility of vanadium(V) ligation to glutathione.  相似文献   

17.
The feed and feces of a continuously fed sheep were analyzed for carbon, hydrogen, and nitrogen, with oxygen as the remainder. The daily feed-feces weight difference was used as the reactant in an equation representing the rumen fermentation. The measured products were the daily production of volatile fatty acids (VFA), CH(4), CO(2), and ammonia. The carbon unaccounted for was assumed to be in the microbial cell material produced in the rumen and absorbed before reaching the feces. The ratio of C to H, O, and N in bacteria was used to represent the elemental composition of the microbes formed in the rumen fermentation, completing the following equation:C(20.03)H(36.99)O(17.406)N(1.345) + 5.65 H(2)O --> C(12)H(24)O(10.1) + 0.83 CH(4) VFA + 2.76 CO(2) + 0.50 NH(3) + C(4.44)H(8.88)O(2.35)N(0.785) microbial cells absorbed With C arbitrarily balanced and O balanced by appropriate addition of water, any error is reflected in the H. The H recovery was 98.5%. The turnover rate constant for rumen liquid equilibrating with polyethylene glycol (PEG) was 2.27 per day. Direct counts and volume measurements of the individual types of bacteria and protozoa in the rumen were used to calculate the total microbial cell volume in the rumen, not equilibrating with it. The dry matter in the rumen (582 g) and the nitrogen content (12.05) of the microbes in the rumen were estimated, the latter constituting 85% of the measured N in the rumen. Calculations for rumen dry matter and nitrogen turning over at the PEG rate introduce big discrepancies with other parameters; a rumination pool must be postulated. Its size and composition are estimated. Arguments are presented to support the view that dry matter and some of the microbes, chiefly the protozoa, do not leave the rumen at the PEG rate. One experiment with the same sheep fed twice daily showed significantly less production of microbial cells than did the continuous (each 2 hr) feeding. Analysis of the microbial cell yield suggests that, on the basis of 11 mg of cells per adenosine triphosphate molecule, a maximum of six adenosine triphosphate molecules could have been formed from each molecule of hexose fermented.  相似文献   

18.
Batch and continuous cultures were used to compare specific physiological features of the hyperthermophilic archaeon, Thermococcus litoralis (T(opt) of 85 degrees to 88 degrees C), to another fermentative hyperthermophile that reduces S degrees facultatively, that is, the bacterium Thermotoga maritima (T(opt) of 80 degrees to 85 degrees C). Under nutritionally optimal conditions, these two hyperthermophiles had similar growth yields on maltose and similar cell formula weights based on elemental analysis: CH(1.7)O(0. 7)N(0.2)S(0.006) for T. litoralis and CH(1.6)O(0.6)N(0.2)S(0.005) for T. maritima. However, they differed with respect to nitrogen source, fermentation product patterns, and propensity to form exopolysaccharides (EPS). T. litoralis could be cultured in the absence or presence of maltose on an amino acid-containing defined medium in which amino acids served as the sole nitrogen source. T. maritima, on the other hand, did not utilize amino acids as carbon, energy, or nitrogen sources, and could be grown in a similar defined medium only when supplemented with maltose and ammonium chloride. Not only was T. litoralis unable to utilize NH(4)Cl as a nitrogen source, its growth was inhibited at certain levels. At 1 g/L ( approximately 20 mM) NH(4)Cl, the maximum growth yield (Y(x/s(max))) for T. litoralis was reduced to 13 g cells dry weight (CDW)/mol glucose from 40 g CDW/mol glucose in media lacking NH(4)Cl. Alanine production increased with increasing NH(4)Cl concentrations and was most pronounced if growth on NH(4)Cl was carried out in an 80% H(2) atmosphere. In T. maritima cultures, which would not grow in an 80% H(2) atmosphere, alanine and EPS were produced at much lower levels, which did not change with NH(4)Cl concentration. EPS production rose sharply at high dilution rates for both organisms, such that maltose utilization plots were biphasic. Wall growth effects were also noted, because cultures failed to wash out at dilution rates significantly above maximum growth rates determined from batch growth experiments. This study illustrates the importance of effective cultivation methods for addressing physiological issues related to the growth of hyperthermophilic heterotrophs.  相似文献   

19.
开放式空气CO2增高对稻田CH4和N2O排放的影响   总被引:12,自引:3,他引:9  
在FACE(free aircarbondioxideenrichment)平台上 ,采用静态暗箱 气相色谱法观测研究了大气CO2 浓度增加对稻田CH4和N2 O排放的影响 .结果表明 ,在 15 0和 2 5 0kgN·hm-2 两种氮肥水平下大气CO2 浓度增加 2 0 0 μmol·mol-1均明显促进水稻生长 ,水稻生物量积累 .大气CO2 浓度增加对 15 0和 2 5 0kgN·hm-2 两种氮肥水平下稻田CH4排放均无显著影响 ,并简要分析了与现有文献报道结果不一致的原因 .大气CO2 浓度增加也未导致 15 0和 2 5 0kgN·hm-2 两种氮肥水平下稻田N2 O排放的明显变化 ,与大多数研究结果一致 .  相似文献   

20.
The aim of this study was to compare the environmental impact of the straw-flow system for fattening pigs with the slatted-floor system by measuring pollutant gas emissions such as ammonia (NH3), nitrous oxide (N2O), methane (CH4) and carbon dioxide (CO2), manure nitrogen (N) content and emissions of water vapour (H2O). Three successive batches of 32 pigs were fattened. For each batch, pigs were allotted to two groups raised in separated rooms fitted either with a concrete totally slatted-floor system (0.75 m2 per pig) or with a straw-flow system (0.79 m2 per pig). With this last system, pigs were kept on a sloped floor, straw being provided daily at the top of the pen. Throughout the fattening period, about 34.4 kg of straw were supplied per pig. The straw, mixed with dung, travelled down the slope by pig motion and went out of the pen to a scraped passage. The solid fraction was scraped every day, stored in a heap in the room and removed every month, 1 week before each period of gaseous emission measurement. The liquid fraction was automatically pumped from the scraped passage into a hermetic tank, which was emptied at the end of each fattening period. Rooms were ventilated mechanically in order to maintain a constant ambient temperature. Once a month, the emissions of NH3, N2O, CH4, CO2 and H2O were measured hourly for 6 consecutive days via infrared photoacoustic detection. Mean daily emissions per pig fattened on the slatted floor or on the sloped floor were, respectively, 4.98 and 13.31 g NH3, 0.67 and 0.68 g N2O, 15.2 and 8.88 g CH4, 548 g and 406 g CO2 equivalents, 1.61 and 1.77 kg CO2 and 2.33 and 2.95 kg H2O. Except for N2O emissions, all the differences were statistically significant (P < 0.001). From the slatted-floor system, the amount of slurry removed per fattening period was on average 256 kg per pig. From the straw-flow system, solid manure amounted on average to 209 kg per pig and liquid manure to 53 kg per pig. The total N-content of the manure was 2.23 kg N per pig with the straw-flow system (solid and liquid manure) v. 3.26 kg N per pig for slurry from the slatted-floor system. This reduction of 30% observed with the sloped floor was mainly explained by the higher level of NH3-N emissions.  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