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.

Background  

Biogenic emissions of methyl halides (CH3Cl, CH3Br and CH3I) are the major source of these compounds in the atmosphere; however, there are few reports about the halide profiles and strengths of these emissions. Halide ion methyltransferase (HMT) and halide/thiol methyltransferase (HTMT) enzymes concerning these emissions have been purified and characterized from several organisms including marine algae, fungi, and higher plants; however, the correlation between emission profiles of methyl halides and the enzymatic properties of HMT/HTMT, and their role in vivo remains unclear.  相似文献   

2.
A monohalomethane-producing enzyme, S-adenosyl-L-methionine-dependent halide ion methyltransferase (EC 2.1.1.-) was purified from the marine microalga Pavlova pinguis by two anion exchange, hydroxyapatite and gel filtration chromatographies. The methyltransferase was a monomeric molecule having a molecular weight of 29,000. The enzyme had an isoelectric point at 5.3, and was optimally active at pH 8.0. The Km for iodide and SAM were 12 mM and 12 microM, respectively, which were measured using a partially purified enzyme. Various metal ions had no significant effect on methyl iodide production, suggesting that the enzyme does not require metal ions. The enzyme reaction strictly depended on SAM as a methyl donor, and the enzyme catalyzed methylation of the I-, Br-, and Cl- to corresponding monohalomethanes and of bisulfide to methyl mercaptan.  相似文献   

3.
5'-[N-[(3S)-3-Amino-carboxypropyl]-N-methylamino]-5(')-deoxyadenosine (azaSAM), an analog of S-adenosyl-L-methionine (SAM), was used to study the cofactor-dependent reduction of the [4Fe-4S](2+) center in lysine 2,3-aminomutase to the +1 oxidation state. azaSAM has a tertiary nitrogen in place of the sulfonium center of SAM. The analog binds to lysine 2,3-aminomutase with K(d)s of 1.4+/-0.3 microM at pH 8.0 and 2.2+/-0.6 microM at pH 6.5. Reduction of the [4Fe-4S](2+) center in the presence of this analog gives a 10K [4Fe-4S](1+) electron paramagnetic resonance (EPR) signal similar to that seen with SAM or S-adenosyl-L-homocysteine (SAH). The pH dependence of cofactor-induced reduction was examined to determine whether ionization of the tertiary nitrogen (pK(a)=7.08) might affect reduction of the [4Fe-4S](2+) center. The results show similar behavior in azaSAM and SAH, demonstrating that ionization of the aza group in azaSAM does not account for pH dependence in cofactor-dependent reduction of the [4Fe-4S](2+) center. The signal shape of the low-temperature EPR signal for the [4Fe-4S](1+) center in the SAM-induced reduction displayed a pH dependence that was not observed in the azaSAM- or SAH-induced spectra. Unique features of the signal are at a maximum at the pH activity optimum of pH 8 and are diminished as the pH is lowered or raised. These features are also absent in the spectra at all pHs examined when reduction is induced by azaSAM or SAH.  相似文献   

4.
陆地生态系统卤甲烷释放特点及其生态意义   总被引:2,自引:0,他引:2  
大气卤甲烷与平流层臭氧破坏密切相关,并参与光化学反应,还具有一定的.温室效应和污染毒害作用。研究发现:(1)大气CH3Cl和CH3Br存在巨大的未知源,它们的已知源分别仅占已知汇的大约1/2~2/3和60%。而CH3I的源和汇还都不确切;(2)陆地生态系统有可能是最大的卤甲烷自然释放源;(3)生物合成和土壤非生物生产是陆地生态系统卤甲烷生产的两个主要途径;(4)沿海湿地、水稻田、热带森林等陆地生态系统是卤甲烷主要释放源;(5)陆地生态系统卤甲烷的自然释放可能在生物竞争、生物代谢和大气环境污染方面具有重要的生态意义;(6)随着大气卤甲烷人为释放源的控制,其自然释放源的相对重要性将更加突出。提出了当前陆地生态系统卤甲烷释放研究的重点方向以及我国开展相关研究的重要意义。  相似文献   

5.
A monohalomethane-producing enzyme, S-adenosyl-L-methionine-dependent halide ion methyltransferase (EC 2.1.1.-) was purified from the marine microalga Pavlova pinguis by two anion exchange, hydroxyapatite and gel filtration chromatographies. The methyltransferase was a monomeric molecule having a molecular weight of 29,000. The enzyme had an isoelectric point at 5.3, and was optimally active at pH 8.0. The Km for iodide and SAM were 12 mM and 12 μM, respectively, which were measured using a partially purified enzyme. Various metal ions had no significant effect on methyl iodide production, suggesting that the enzyme does not require metal ions. The enzyme reaction strictly depended on SAM as a methyl donor, and the enzyme catalyzed methylation of the I-,Br-, and Cl- to corresponding monohalomethanes and of bisulfide to methyl mercaptan.  相似文献   

6.
Methyl chloride (CH(3)Cl) and methyl bromide (CH(3)Br) are the primary carriers of natural chlorine and bromine, respectively, to the stratosphere, where they catalyze the destruction of ozone, whereas methyl iodide (CH(3)I) influences aerosol formation and ozone loss in the boundary layer. CH(3)Br is also an agricultural pesticide whose use is regulated by international agreement. Despite the economic and environmental importance of these methyl halides, their natural sources and biological production mechanisms are poorly understood. Besides CH(3)Br fumigation, important sources include oceans, biomass burning, tropical plants, salt marshes, and certain crops and fungi. Here, we demonstrate that the model plant Arabidopsis thaliana produces and emits methyl halides and that the enzyme primarily responsible for the production is encoded by the HARMLESS TO OZONE LAYER (HOL) gene. The encoded protein belongs to a group of methyltransferases capable of catalyzing the S-adenosyl-L-methionine (SAM)-dependent methylation of chloride (Cl(-)), bromide (Br(-)), and iodide (I(-)) to produce methyl halides. In mutant plants with the HOL gene disrupted, methyl halide production is largely eliminated. A phylogenetic analysis with the HOL gene suggests that the ability to produce methyl halides is widespread among vascular plants. This approach provides a genetic basis for understanding and predicting patterns of methyl halide production by plants.  相似文献   

7.
Despa S  Bers DM 《Biophysical journal》2003,84(6):4157-4166
Na/K pump current (I(pump)) and intracellular Na concentration ([Na](i)) were measured simultaneously in voltage-clamped rabbit ventricular myocytes, under conditions where [Na](i) is controlled mainly by membrane transport. Upon abrupt pump reactivation (after 10-12 min blockade), I(pump) decays in two phases. Initially, I(pump) declines with little [Na](i) change, whereas the second phase is accompanied by [Na](i) decline. Initial I(pump) sag was still present at external [K] = 15 mM, but prevented by [Na](i) approximately 100 mM. Initial I(pump) sag might be explained by subsarcolemmal [Na](i) ([Na](SL)) depletion produced by rapid Na extrusion and I(pump). Brief episodes of pump blockade allowed [Na](SL) repletion, since peak postblockade I(pump) exceeded I(pump) at the end of previous activation (without appreciably altered global [Na](i)). The apparent K(m) for [Na](i) was higher for continuous I(pump) activation than peak I(pump) (14.1 +/- 0.2 vs. 11.2 +/- 0.2 mM), whereas that based on d[Na](i)/dt matched peak I(pump) (11.6 +/- 0.3 mM). [Na](SL) depletion (vs. [Na](i)) could be as high as 3 mM for [Na](i) approximately 18-20 mM. A simple diffusion model indicates that such [Na](SL) depletion requires a Na diffusion coefficient 10(3)- to 10(4)-fold below that expected in bulk cytoplasm (although this could be subsarcolemmal only). I(pump) integrals and [Na](i) decline were used to estimate intracellular Na buffering, which is slight (1.39 +/- 0.09).  相似文献   

8.
N-Methyltransferase catalyzing the transfer of methyl groups from [14CH3]SAM to xanthosine, producing 7- methylxanthosine, was extracted in cell-free systems from tea leaves. The purine nucleotides, nucleosides, and bases, except for xanthosine, were all inactive substrates in the methylation at the N-7 position of their purine rings. This confirmed that the methylation of xanthosine to 7- methylxanthosine is the first step of methylation in the pathway for caffeine biosynthesis. The optimum pH for the methyltransferase is 7.5 to 8.0. PCMB (0.5 mM), Zn2 + (1 mM) and Cu2+ (1 mM) strongly inhibited the enzyme activity. The Km values for xanthosine and SAM are 0.25 mM at 3.9µM of SAM and 3.3µM at 0.40 mM of xanthosine, respectively.  相似文献   

9.
The steady-state kinetic properties of partially purified chicken liver sorbitol dehydrogenase (SDH) were determined spectrophotometrically at 25 degrees C, in 50 mM 3-(N-morpholino)propanesulfonic acid (MOPS) buffer, pH 8.0. In the sorbitol-to-fructose direction, analysis was based on initial rate data obtained at [NAD(+)](o)=0.1-0.4 mM and [sorbitol](o)=1.25-10 mM. The reverse process was analyzed by recording progress curves for NADH consumption, starting with [NADH](o)=0.2 mM and [fructose](o)=66.7-267 mM. The kinetics conformed to an ordered sequential model, with the cofactors adding first. The steady-state parameters in the forward direction, K(NAD(+)), K(iNAD(+)) and K(sorbitol), were found to be 210+/-62 muM, 220+/-69 microM and 3.2+/-0.54 mM, respectively. The corresponding parameters in the reverse direction were K(NADH)=240+/-58 microM, K(iNADH)=10+/-2.8 microM and K(fructose)=1000+/-140 mM. The results indicated a close parallelism with human SDH, yet up to 40-fold differences were observed when compared to related reports on other mammalian species. The structural and adaptive bases of the variation in substrate and cofactor affinities need to be accounted for.  相似文献   

10.
A new Ru(II) complex, [Ru(bpy)(2)(dhipH3)](ClO4)(2) (in which bpy=2,2'-bipyridine, dhipH(3)=3,4-dihydroxy-imidado[4,5-f][1,10]-phenanthroline), was synthesized and characterized, and the pH effect on the emission spectra of the complex was studied. The interaction of the complex with calf thymus DNA was investigated by UV-visible and emission spectroscopy, and viscosity measurements. The results suggest that the complex acted as a sensitive luminescent pH sensor and a strong ct-DNA intercalator with an intrinsic binding constant of (4.0+/-0.7) x 10(5) M(-1) in buffered 50 mM NaCl.  相似文献   

11.
Two new ruthenium(II) complexes of [Ru(bpy)(pp[2,3]p)2](ClO4)2 and [Ru(phen)(pp[2,3]p)2](ClO4)(2) (bpy=2,2'-bipyridine, phen=1,10-phenanthroline, pp[2,3]p=pyrido[2',3':5,6]pyrazino[2,3-f][1,10]phenanthroline) have been synthesized and characterized by elemental analysis and 1H NMR spectra. The calf thymus DNA-binding properties of the two complexes were investigated by UV-visible and emission spectroscopy, competitive binding experiments with ethidium bromide and viscosity measurements. The results indicate that the two complexes intercalate between the base pairs of the DNA tightly with intrinsic DNA-binding constants of 3.08 x 10(6) and 6.53 x 10(6) M(-1) in buffered 50 mM NaCl, respectively, which are much larger than 6.9 x 10(5) M(-1) for [Ru(bpy)2(pp[2,3]p)](ClO4)2 containing two ancillary ligands of bpy.  相似文献   

12.
It was found that all halides can compete with cyanide for binding with myeloperoxidase. The lower is the pH, the higher is the affinity of halides. The apparent dissociation constants (Kd) of myeloperoxidase-cyanide complex were determined in the presence of F-, Cl-, Br- and I- in the pH range of 4 to 7. In slightly acidic pH (4 - 6) fluoride and chloride exhibit a higher affinity towards the enzyme than bromide and iodide. Taking into account competition between cyanide and halides for binding with myeloperoxidase the dissociation constants of halide-myeloperoxidase complexes were calculated. All halides except fluoride can be oxidized by H2O2 in the presence of myeloperoxidase. However, since fluoride can bind with myeloperoxidase, it can competitively inhibit the oxidation of other halides. Fluoride was a competitive inhibitor with respect to other halides as well as to H2O2. Inhibition constants (Ki) for fluoride as a competitive inhibitor with respect to H2O2 increased from iodide oxidation through bromide to chloride oxidation.  相似文献   

13.
The thyroid gland accumulates iodide for the synthesis of thyroid hormones. The aim of the current study was to quantify iodide accumulation in cultured thyroid cells by live cell imaging using the halide-sensitive yellow fluorescent protein (YFP) variant YFP-H148Q/I152L. In vivo calibrations were performed in FRTL-5 thyrocytes to determine the sensitivity of YFP-H148Q/I152L to iodide. In the presence of ion-selective ionophores, YFP-H148Q/I152L fluorescence was suppressed by halides in a pH-dependent manner with 20-fold selectivity for iodide versus chloride and competition between the two halides. At a physiological pH of 7 and a chloride concentration of 15mM, the affinity constant of YFP-H148Q/I152L for iodide was 3.5mM. In intact FRTL-5 cells, iodide induced a reversible decrease in YFP-H148Q/I152L fluorescence. FRTL-5 cells concentrated iodide to 60 times the extracellular concentration. Iodide influx exhibited saturation kinetics with respect to extracellular iodide with a K(m) of 35 microM and a V(max) of 55 microM/s. Iodide efflux exhibited saturation kinetics with respect to intracellular iodide concentration with a K(m) of 2.2mM and a V(max) of 43 microM/s. The results of this study demonstrate the utility of YFP-H148Q/I152L as a sensitive and selective biosensor for the quantification of iodide accumulation in thyroid cells.  相似文献   

14.
The [Ru(II)(Hedta)NO(+)] complex is a diamagnetic species crystallizing in a distorted octahedral geometry, with the Ru-N(O) length 1.756(4) A and the RuNO angle 172.3(4) degrees . The complex contains one protonated carboxylate (pK(a)=2.7+/-0.1). The [Ru(II)(Hedta)NO(+)] complex undergoes a nitrosyl-centered one-electron reduction (chemical or electrochemical), with E(NO+/NO)=-0.31 V vs SCE (I=0.2 M, pH 1), yielding [Ru(II)(Hedta)NO](-), which aquates slowly: k(-NO)=2.1+/-0.4x10(-3) s(-1) (pH 1.0, I=0.2 M, CF(3)COOH/NaCF(3)COO, 25 degrees C). At pHs>12, the predominant species, [Ru(II)(edta)NO](-), reacts according to [Ru(II)(edta)NO](-)+2OH(-)-->[Ru(II)(edta)NO(2)](3-), with K(eq)=1.0+/-0.4 x 10(3) M(-2) (I=1.0 M, NaCl; T=25.0+/-0.1 degrees C). The rate-law is first order in each of the reactants for most reaction conditions, with k(OH(-))=4.35+/-0.02 M(-1)s(-1) (25.0 degrees C), assignable mechanistically to the elementary step comprising the attack of one OH(-) on [Ru(II)(edta)NO](-), with subsequent fast deprotonation of the [Ru(II)(edta)NO(2)H](2-) intermediate. The activation parameters were DeltaH(#)=60+/-1 kJ/mol, DeltaS(#)=-31+/-3 J/Kmol, consistent with a nucleophilic addition process between likely charged ions. In the toxicity up-and-down tests performed with Swiss mice, no death was observed in all the doses administered (3-9.08 x 10(-5) mol/kg). The biodistribution tests performed with Wistar male rats showed metal in the liver, kidney, urine and plasma. Eight hours after the injection no metal was detected in the samples. The vasodilator effect of [Ru(II)(edta)NO](-) was studied in aortic rings without endothelium, and was compared with sodium nitroprusside (SNP). The times of maximal effects of [Ru(II)(edta)NO](-) and SNP were 2 h and 12 min, respectively, suggesting that [Ru(II)(edta)NO](-) releases NO slowly to the medium in comparison with SNP.  相似文献   

15.
A novel binuclear complex [(bpy)2Ru(mu-bipp)Ru(bpy)2](ClO4)4, where bpy=2,2'-bipyridine and bipp=2,9-bis(2-imidazo[4,5-f][1,10]phenanthroline)-1,10-phenanthroline has been synthesized. Photophysical results reveal that this complex interacts with calf-thymus DNA with intrinsic binding constant 2.6 x 10(5) M(-1) in the buffer containing 5 mM Tris and 50 mM NaCl. The fact that the intraligand transition of bipp around 370 nm decreased by up to 50% in the presence of DNA, much more pronounced than the metal to ligand charge transfer band around 445 nm indicates the bridging ligand bipp is also the intercalating ligand into DNA base pairs. The emission band around at 601 nm increased by 1.4-fold, and red shifted 14 nm when DNA was added to saturation. The emission quenching of this complex by K4[Fe(CN)6] was depressed greatly when DNA was present. Viscometric measurements also proved the intercalative binding mode.  相似文献   

16.
The perfluoroalkyl compounds (PFCs), perfluoroalkyl sulfonates (PFXS) and perfluoroalkyl carboxylates (PFXA) are environmentally persistent and recalcitrant towards most conventional water treatment technologies. Here, we complete an in depth examination of the UV-254 nm production of aquated electrons during iodide photolysis for the reductive defluorination of six aquated perfluoroalkyl compounds (PFCs) of various headgroup and perfluorocarbon tail length. Cyclic voltammograms (CV) show that a potential of +2.0 V (vs. NHE) is required to induce PFC oxidation and -1.0 V is required to induce PFC reduction indicating that PFC reduction is the thermodynamically preferred process. However, PFCs are observed to degrade faster during UV(254 nm)/persulfate (S(2)O(8)(2-)) photolysis yielding sulfate radicals (E° = +2.4 V) as compared to UV(254 nm)/iodide (I(-)) photolysis yielding aquated electrons (E° = -2.9 V). Aquated electron scavenging by photoproduced triiodide (I(3)(-)), which achieved a steady-state concentration proportional to [PFOS](0), reduces the efficacy of the UV/iodide system towards PFC degradation. PFC photoreduction kinetics are observed to be dependent on PFC headgroup, perfluorocarbon chain length, initial PFC concentration, and iodide concentration. From 2 to 12, pH had no observable effect on PFC photoreduction kinetics, suggesting that the aquated electron was the predominant reductant with negligible contribution from the H-atom. A large number of gaseous fluorocarbon intermediates were semi-quantitatively identified and determined to account for ~25% of the initial PFOS carbon and fluorine. Reaction mechanisms that are consistent with kinetic observations are discussed.  相似文献   

17.
Ectomycorrhizal fungi: A new source of atmospheric methyl halides?   总被引:2,自引:1,他引:1  
Incomplete source budgets for methyl halides – compounds that release inorganic chlorine and bromine radicals which, in turn, catalyze atmospheric ozone depletion – limit our ability to predict the fate of the stratospheric ozone layer. We report here the first measured emissions of methyl chloride, methyl bromide, and methyl iodide from ectomycorrhizal fungi. We grew nine fungal isolates on growth media containing halide concentrations similar to those found in soils and plant tissues. The observed range of emissions was 0.003–65 μg methyl chloride, 0.001–3 μg methyl bromide, and 0.02–12 μg methyl iodide g?1 dry weight fungi day?1. Species varied in production rates of methyl chloride vs. methyl bromide vs. methyl iodide. Cenococcum geophilum, a widespread ectomycorrhizal fungus, was further tested to investigate the effects of halide substrate concentration in growth media. Emissions from this species increased linearly with increasing concentrations of both bromide and iodide. In addition, a subset of four fungi was studied with two media concentrations each of chloride, bromide, and iodide (0.2 or 20 mm ). These fungi had similar responses to halide concentration, despite 1000‐fold differences in baseline emission rates between isolates. Finally, high chloride concentrations (20 mm ) in media did not appear to inhibit emissions of methyl bromide or methyl iodide. Overall, ectomycorrhizal fungi might be an important source of methyl halides to the atmosphere, and substrate concentrations and community composition may influence production levels in ecosystems.  相似文献   

18.
Binding of an iodide ion to horseradish peroxidase was studied by following the hyperfine-shifted proton nuclear magnetic resonance signals of the enzyme. For the enzyme in an iodide-free solution, the spectra of hyperfine-shifted methyl region were only slightly affected by varying pH. In the presence of iodide (200 mM), however, both chemical shifts and line widths of the heme peripheral 1- and 8-methyl proton signals were markedly affected by the pH change from 7 to 4 and broadened at pH 4. From the change in peak heights of these signals at various concentrations of iodide, the dissociation constant of the iodide to the enzyme was calculated to be about 100 mM at pH 4.0. The peak derived from the proximal histidyl imidazole N epsilon-H proton was not perturbed by the addition of 200 mM iodide at pH 4.0 and 7.1. The rate of oxidation of iodide with hydrogen peroxide catalyzed by the enzyme was increased with decreasing pH, indicating the participation of an ionizable group with the pKa value of 4.0. Optical difference spectrum studies showed that iodide exerts no effect both at pH 4.0 and 7.4 on the binding affinity of resorcinol which is associated with the enzyme in the vicinity of the heme peripheral 8-CH3 group. These results suggest that an iodide ion binds to the enzyme at almost equal distance from the heme peripheral 1- and 8-methyl groups at the distal side of the heme and that the interaction becomes stronger in acidic medium with protonation of the ionizable group with the pKa value of 4.0.  相似文献   

19.
Via the Na/Ca and Na/H exchange, intracellular Na concentration ([Na](i)) is important in regulating cardiac Ca and contractility. Functional data suggest that [Na](i) might be heterogeneous in myocytes that are not in steady state, but little direct spatial information is available. Here we used two-photon microscopy of SBFI to spatially resolve [Na](i) in rat ventricular myocytes. In vivo calibration yielded an apparent K(d) of 27 +/- 2 mM Na. Similar resting [Na](i) was found using two-photon or single-photon ratiometric measurements with SBFI (10.8 +/- 0.7 vs. 11.1 +/- 0.7 mM). To assess longitudinal [Na](i) gradients, Na/K pumps were blocked at one end of the myocyte (locally pipette-applied K-free extracellular solution) and active in the rest of the cell. This led to a marked increase in [Na](i) at sites downstream of the pipette (where Na enters the myocyte and Na/K pumps are blocked). [Na](i) rise was smaller at upstream sites. This resulted in sustained [Na](i) gradients (up to approximately 17 mM/120 microm cell length). This implies that Na diffusion in cardiac myocytes is slow with respect to trans-sarcolemmal Na transport rates, although the mechanisms responsible are unclear. A simple diffusion model indicated that such gradients require a Na diffusion coefficient of 10-12 microm(2)/s, significantly lower than in aqueous solutions.  相似文献   

20.
Poly(lactic acid) (PLA) and poly(acrylic acid) (PAA) biomaterials with luminescent ruthenium tris(bipyridine) centers couple drug delivery and imaging functions. Hydrophobic [Ru(bpyPLA2)3](PF6)2 (1) was generated from [Ru[bpy(CH2OH)2]3](PF6)2 in bulk monomer using 4-(dimethylamino)pyridine as the catalyst. The bromoesters, [Ru[bpy(CH2OR)2]3](PF6)2, [Ru[bpy(C13H27)2][bpy(CH2OR]2](PF6)2 (4), and [Ru[bpy(PLAOR)2]3]2+ (9) (R=COCBr(CH3)2), served as initiators for tert-butyl acrylate (tBA) polymerization. Conversion of PtBA to PAA via hydrolysis affords water soluble materials, [Ru(bpyPAA2)3]2+ (7) and [Ru[bpy(C13H27)2](bpyPAA2)2]2+ (8) and the amphiphilic star polymer [Ru[bpy(PLA-PAA)2]3)](PF6)2 (11), which is soluble in a H2O/CH3CN (1:1) mixture. Luminescence excitation and emission spectra of the Ru polymers were in agreement with the parent [Ru(bpy)3]2+ chromophore (lambdaex=468, lambdaem=621 nm). Lifetimes of tau approximately 700 ns in both air and nitrogen atmospheres are typical for most materials; however, the amphiphilic star block copolymer 11 is quenched by oxygen to some degree. Thermal analysis shows the expected glass transitions for the polymeric ruthenium complex materials.  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