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.
Sulfite reductase (SiR) catalyzes a six electron and six proton reduction of sulfite to sulfide. Similarly to the cytochrome P450 (cytP450) family, the active site in SiR contains a (partially reduced) heme bound axially to a cysteinate ligand—though with an extra Fe4S4 cluster. Fe(III) SO2−, Fe(III) SOH, and Fe(III) SO(H2) intermediates have been proposed for the catalytic cycle of SiR, leading to a formally Fe(V)S species—akin to the widely accepted reaction mechanism in cytP450. Here, density functional theory (DFT) data is reported for of such FeSO(H2) intermediates. The Fe(III) SO2− models display relatively high energies for homolytic bond breaking compared to their isomeric oxygen‐bound Fe(III) OS2− models, and thus offer a better alternative in terms of avoiding radical side products able to induce enzyme suicide. This could be due to the fact that the (iron‐bound) sulfur is more active from a redox standpoint compared to oxygen, thus permitting the departing oxygen to maintain a redox‐inert state. Di‐protonation of the oxygen is computed to lead to a compound I type Fe(IV)S coupled to a porphyrin radical anion—consistent with an intermediate previously observed by x‐ray crystallography.  相似文献   

2.
Sulfur poisoning and regeneration are global challenges for metal catalysts even at the ppm level. The sulfur poisoning of single-metal-site catalysts and their regeneration is worthy of further study. Herein, sulfur poisoning and self-recovery are first presented on an industrialized single-Rh-site catalyst (Rh1/POPs). A decreased turnover frequency of Rh1/POPs from 4317 h−1 to 318 h−1 was observed in a 1000 ppm H2S co-feed for ethylene hydroformylation, but it self-recovered to 4527 h−1 after withdrawal of H2S, whereas the rhodium nanoparticles demonstrated poor activity and self-recovery ability. H2S reduced the charge density of the single Rh atom and lowered its Gibbs free energy with the formation of inactive (SH)Rh(CO)(PPh3-frame)2, which could be regenerated to active HRh(CO)(PPh3-frame)2 after withdrawing H2S. The mechanism and the sulfur-related structure–activity relationship were highlighted. This work provides an understanding of heterogeneous ethylene hydroformylation and sulfur-poisoned regeneration in the science of single-atom catalysts.  相似文献   

3.
Biomass gasification using supercritical water is a promising way to produce hydrogen gas. However, this method might release toxic heteroatomic compounds. It is therefore important to clarify reaction pathways for efficiently obtaining hydrogen gas and suppressing environmental burden. L-cysteine was adopted for a test reagent containing sulfur and determination of the sulfur compound reaction pathways was studied by Li+-ion attachment mass spectrometry. It was found that H2S, CO, CO2, SO, SO2 and SO3 gasses were released at high concentrations in the gas phase during the hydrothermal reaction. By adding Ca(OH)2 as alkali, the pathway of these gasses were, however, suppressed into the liquid phase so that the toxic emissions to the gas phase could be avoided.  相似文献   

4.
《Analytical letters》2012,45(9):1333-1345
Abstract

An operationally inexpensive and satisfactory analytical procedure for sulfur dioxide is proposed. The reagent 3-methyl-1,2-cyclopentanodione dithiosemicarbazone has been used to determine trace amounts of sulfur dioxide indirectly using the reduction of Fe(III) to Fe(II) principle. In order to find the optimal conditions for SO2 determination, properties of 3-Me-CPDT-Fe(II) complex such as its composition, stability and free energy change of formation have been determined. The best conditions for the complex formation such as standing time, pH, wavelength and the effect of interfering ions are described. The complex has been used with success in spectrophotometric determination of SO2. The procedure can determine down to 0.63 μg and recoveries are better than 97.5%. The method is also suitable for determination of sulfur dioxide in the air provided that interfering gases such as H2S and NO2 are removed.  相似文献   

5.
The generation of negative ions from SO2 in the gas phase was studied using the thermal surface ionization method. Six anion types were measured: O, S, SO, and SO2 and anions with m/z=96 and m/z=128. The most abundant anion formed was S and the formation routes are discussed for each of the six anions. O, S, and SO are formed via dissociative electron attachment to the molecule, whereas the generation of SO2 and anions with m/z=96 and m/z=128 are probably associated with the formation of H2SO4 in the gas inlet system and the ion source. Using statistical thermodynamics the dissociation temperatures of SO2 and SO in the gas phase are calculated and values of above 1800 °C are obtained for both molecules. We also estimated the optimal filament temperatures for the formation of all anions measured, indicating that for SO2 the optimal temperature is related to the electron affinity of the molecules: the optimal temperature increases with decreasing value of the electron affinity for the molecule corresponding to the respective anion.  相似文献   

6.
The iodine–sulfur (IS) thermochemical process for hydrogen production is one of the most promising approaches in using high‐temperature process heat supplied by a nuclear reactor. This process includes three reactions that form a closed cycle: the Bunsen reaction, in which iodine, water, and sulfur dioxide react to form sulfuric acid and hydriodic acid (HI); HI decomposition; and sulfuric acid decomposition. However, the side reactions between H2SO4 and HI may disturb the operation of the IS closed cycle. For optimal process conditions, the reaction kinetics between H2SO4 and HI should be examined. In this work, a preliminary kinetic study was conducted. Using the initial reaction rate method, the kinetic parameters of the reaction between sulfuric acid and HI, such as the apparent reaction orders and rate constant were determined. For I?, the apparent reaction order was approximately 1.77, whereas the orders for H+ and SO42? were 7.78 and 1.29, respectively. The apparent rate constant at 85 ± 1°C was approximately 2.949 × 10?11 min?1 (mol/L)?9.84. The H+ concentration had more significant influence on the reaction rate than those of SO42? and I?. Such basic data provide useful information for related process design and further kinetics study.  相似文献   

7.
Carbon materials slightly doped with heteroatoms such as nitrogen (N-RFC) or sulfur (S-RFC) are investigated as active catalysts for the electrochemical bielectronic oxygen reduction reaction (ORR) to H2O2. Mesoporous carbons with wide, accessible pores were prepared by pyrolysis of a resorcinol-formaldehyde resin using a PEO-b-PS block copolymer as a sacrificial templating agent and the nitrogen and sulfur doping were accomplished in a second thermal treatment employing 1,10-phenanthroline and dibenzothiophene as nitrogen and sulfur precursors, respectively. The synthetic strategy allowed to obtain carbon materials with very high surface area and mesopore volume without any further physicochemical post treatment. Voltammetric rotating ring-disk measurements in combination with potentiostatic and galvanostatic bulk electrolysis measurements in 0.5 m H2SO4 demonstrated a pronounced effect of heteroatom doping and mesopores volume on the catalytic activity and selectivity for H2O2. N-RFC electrode was employed as electrode material in a 45 h electrolysis showing a constant H2O2 production of 298 mmol g−1 h−1 (millimoles of H2O2 divided by mass of catalyst and electrolysis time), with a faradic efficiency (FE) up to 61 % and without any clear evidence of degradation. The undoped carbon RFC showed a lower production rate (218 mmol g−1 h−1) but a higher FE of 76 %, while the performances drastically dropped when S-RFC (production rate 11 mmol g−1 h−1 and FE=39 %) was used.  相似文献   

8.
An ion chromatography‐inductively coupled plasma mass spectrometric (IC‐ICP‐MS) method for the speciation of sulfur compounds, namely sulfite [SO32?], sulfate [SO42?] and thiosulfate [S2O32?], was described. Ionic sulfur compounds were well separated in about 3 min by ion chromatography with a Hamilton PRP‐X100 column as the stationary phase and 60 mmol L?1 NH4NO3 and 0.1% v/v formaldehyde (HCHO) solution (pH = 7) as the mobile phase. The analyses were carried out using dynamic reaction cell (DRC) ICP‐MS. The sulfur‐selective chromatogram was determined at m/z 48 as 32S16O+ by using its reaction with O2 in the reaction cell. The method avoided the effect of polyatomic isobaric interferences at m/z 32 caused by 16O16O+ and 14N18O+ on 32S+ by detecting 32S+ as the oxide ion 32S16O+ at m/z 48, which is less interfered. The detection limit of various species studied was in the range of 3.6–4.6 ng S mL?1. The accuracy of the method has been verified by comparing the sum of the concentrations of individual sulfur compounds obtained by the present procedure with the total concentration of sulfur in several natural water samples. The recovery was in the range of 97–102% for various compounds studied.  相似文献   

9.
X-ray fluorescence (XRF) spectra of poly(butyl cyanoacrylate) initiated, in tetrahydrofuran, by triphenylphospine and terminated by SO3, show the persistent presence of phosphorus and sulfur in the polymer, in near-equivalent proportions, with the sulfur content partially removable by anion-exchange. It is concluded that all chains have initial phosphonium groups, and, in abesence of any hydrolysis-formed H2SO4, are macrozwitterions with terminal  SO3. Acid-terminated chains should have proton-capped ends and initial phosphonium+ · HSO4 ion-pairs.  相似文献   

10.
Donor molecules undergo dramatic changes in their chemical properties on coordination to a transition-metal atom. Highly reactive species can be trapped and studied as ligands. Conversely, stable compounds can be activated to undergo novel reactions. Sulfur dioxide complexes have generally been studied from a structural viewpoint, their reactivity having been somewhat neglected. The unstable sulfur oxides SO, S2O, and S2O2 are still often regarded as laboratory curiosities. Their successful stabilization in transition-metal complexes has now made them accessible to detailed study, in the course of which many relationships to the chemistry of SO2 complexes have become apparent.  相似文献   

11.
The reduction of cupric ion in ammonia solution by aqueous sulfur dioxide was studied. Each run was carried out at constant initial cupric concentration, stirred rate and total mixed gas flow rate. The effect of temperature, partial pressure of sulfur dioxide in gas phase and cupric ion concentration of the solution was investigated. The reaction of Cu3++SO2(aq.)→Cu++SO42? was carried out by bubbling mixed gas (SO2/N2) through the aqueous ammonia complex of copper (II). The color change for the system was from deep blue, green, yellow to white. The pH values in the system changed from 10 to 4. The product of the reaction was identified by the analyses of IR spectrum and X-ray diffraction, having the formula of 7Cu2SO3· 2CuSO3·3(NH4)2SO3·24H2O. The kinetic model of the reduction was proposed as: –d[Cu2+]/dt = k exp(–E/RT)[Cu2+]α[SO2%] According to the experiments, the parameters were determined as: α=1.64±0.03, þ=1.20, E=13.7 Kcal/mol and k = (1.77±0.20)×1010 (g-equ./?)?0.64min?2.  相似文献   

12.
Although integral to remote marine atmospheric sulfur chemistry, the reaction between methylsulfinyl radical (CH3SO) and ozone poses challenges to theoretical treatments. The lone theoretical study on this reaction reported an unphysically large barrier of 66 kcal mol−1 for abstraction of an oxygen atom from O3 by CH3SO. Herein, we demonstrate that this result stems from improper use of MP2 with a single-reference, unrestricted Hartree-Fock (UHF) wavefunction. We characterized the potential energy surface using density functional theory (DFT), as well as multireference methodologies employing a complete active-space self-consistent field (CASSCF) reference. Our DFT PES shows, in contrast to previous work, that the reaction proceeds by forming an addition adduct [CH3S(O3)O] in a deep potential well of 37 kcal mol−1. An O−O bond of this adduct dissociates via a flat, low barrier of 1 kcal mol−1 to give CH3SO2+O2. The multireference computations show that the initial addition of CH3SO+O3 is barrierless. These results provide a more physically intuitive and accurate picture of this reaction than the previous theoretical study. In addition, our results imply that the CH3SO2 formed in this reaction can readily decompose to give SO2 as a major product, in alignment with the literature on CH3SO reactions.  相似文献   

13.
The protonated species [Fe2(η-C5H5)2(CO)2(η-CO){μ-CN(Me)H}]X, [Fe2(η-C5H5)2(CO)(CNMe)(μ-CO){μ-CN(Me)H}][X], and [Fe2(η-C5H5)2(CO)2{η-CN(Me)H}2][X]2 react with one equivalent of AgY. The Ag+ and one H+ act together as a two-electron oxidant. Silver metal is precipitated quantitatively and the substrates cleaved to give mono-nuclear products of the type (a) [Fe(η-C5H5)(CO)(L)X] and [Fe(η-C5H5(CO)(L)Y] or (b) Fe(η-C5H5(CO)(L)(CNMe)][X] (L = CO, CNMe). If X and Y are both coordinating anions such as NO3, I, or Br or the solvent is MeCN products of type (a) are usually obtained with X = Y = MeCN+ if acetonitrile is used as the solvent. However, if either X or Y is a non-coordinating anion such as BF4 or PF6 and methanol is the solvent, the products are usually those of type (b). When X = [p-MeC6H4SO3], both types of products are obtained in significant amounts. If two equivalents of Ph3P are added to the methanol solution of [Fe2(η-C5H5)2(CO)2{-CN(Me)H}2[BF6]2, no reaction takes place until the third equivalent of AgNO3 has been added. The products have been isolated and characterized by analysis and infrared spectroscopy. The previously unreported [Fe2(η-C5H5)2(CO)(CNMe)(η-CO){η-CN(Me)H}] X salts are described for X = BF4, PF6, Br · 2H2O, I · H2O, NO3 · 0.5H2O, and p-MeC6H4SO3.  相似文献   

14.
(Per)thionitrite (SNO/SSNO) intermediates play vital roles in modulating nitric oxide (NO) and hydrogen sulfide (H2S) dependent bio-signalling processes. Whilst the previous preparations of such intermediates involved reactive H2S/HS or sulfane sulfur (S0) species, the present report reveals that relatively stable thiocarbonyl compounds (such as carbon disulfide (CS2), thiocarbamate, thioacetic acid, and thioacetate) react with nitrite anion to yield SNO/SSNO. For instance, the reaction of CS2 and nitrite anion (NO2) under ambient condition affords CO2 and SNO/SSNO. A detailed investigation involving UV/Vis, FTIR, HRMS, and multinuclear NMR studies confirm the formation of SNO/SSNO, which are proposed to form through an initial nucleophilic attack by nitrite anion followed by a transnitrosation step. Notably, reactions of CS2 and nitrite in the presence of thiol RSH show the formation of organic polysulfides R-Sn-R, thereby illustrating that the thiocarbonyls are capable of influencing the pool of bioavailable sulfane sulfurs. Furthermore, the availability of both NO2 and thiocarbonyl motifs in the biological context hints at their synergistic metal-free activations leading to the generation of NO gas and various reactive sulfur species via SNO/SSNO.  相似文献   

15.
Preparation and Crystal Structure of the Tetramethylammonium Thiocyanate Sulfur Dioxide Adduct, (CH3)4N+SCN · SO2 Tetramethylammonium thiocyanate reacts with sulfur dioxide under formation of tetramethylammonium thiocyanate sulfur dioxide adduct. The resulting salt is characterised by NMR and vibrational spectroscopy and its crystal structure. (CH3)4N+SCN · SO2 crystallizes in the monoclinic space group P21/c with a = 578.4(1) pm, b = 1634.3(1) pm, c = 1054.6(1) pm, β = 105.17(1)°, and four formula units in the unit cell. The crystal structure possesses a strong S–S interaction between the NCS anion and the SO2 molecule. The NCS–SO2 distance of 301.02(9) pm is longer than a covalent single bond, thus the compound is rather described as an adduct. The structure is compared with ab initio calculated data.  相似文献   

16.
The exothermic H-atom abstraction reaction of SO+2 with H2 has been studied in a selected ion flow drift tube (SIFDT) over a range of center-of-mass energies from thermal (300 K) to about 0.12 eV. The measured rate coefficient at 300 K is 4.2 × 10−12 cm3 s−1 which is very much less than the Langevin capture rate. The increase in rate coefficient with ion kinetic energy gives a linear Arrhenius-type plot with a slope that indicates a barrier of ∼5 kJ mol−1 exists on the potential surface. The H2SO+2 potential surface is also explored in an ab initio investigation using the G2 procedure. An (SO+2.H2)1 transition state between reactants and products is identified, corresponding to the barrier found from experiments.  相似文献   

17.
High interconversion energy barriers, depressive reaction kinetics of sulfur species, and sluggish Li+ transport inhibit the wide development of high-energy-density lithium sulfur (Li−S) batteries. Herein, differing from random mixture of selected catalysts, the composite catalyst with outer delocalized isoelectronic heterostructure (DIHC) is proposed and optimized, enhancing the catalytic efficiency for decreasing related energy barriers. As a proof-of-content, the FeCoOxSy composites with different degrees of sulfurization are fabricated by regulating atoms ratio between O and S. The relationship of catalytic efficiency and principal mechanism in DIHCs are deeply understood from electrochemical experiments to in situ/operando spectral spectroscopies i.e., Raman, XRD and UV/Vis. Consequently, the polysulfide conversion and Li2S precipitation/dissolution experiments strongly demonstrate the volcano-like catalytic efficiency of various DIHCs. Furthermore, the FeCoOxSy-decorated cell delivers the high performance (1413 mAh g−1 at 0.1 A g−1). Under the low electrolyte/sulfur ratio, the high loading cell stabilizes the areal capacity of 6.67 mAh cm−2 at 0.2 A g−1. Impressively, even resting for about 17 days for possible polysulfide shuttling, the high-mass-loading FeCoOxSy-decorated cell stabilizes the same capacity, showing the practical application of the DIHCs in improving catalytic efficiency and reaching high electrochemical performance.  相似文献   

18.
Rate constants and product branching ratios were measured for eleven sulfur oxide, sulfur fluoride, and sulfur oxyfluoride anions reacting with O3. The SO 2 ion reacts rapidly to form –O 3, SO 3, and e. The temperature dependence of the branching ratio shows more reactive detachment and less SO 3 formation at higher temperature. SO 3 reacts with O3, forming SO 4 at 1/3 to 1/4 of the collisional rate from 200 to 500 K, respectively. At 300 K, SF 6 charge transfers to O3 at 20% of the collisional rate. F2SO 2 reacts with O3 at a few percent of the collision rate, forming both O 3 and FSO 3; The ion F3SO reacts slowly with O3 to form F3SO 2. The ions SO 4, SF 5, FSO 2, FSO 3, F3SO, and F5SO are unreactive with O3. A trend is noted relating the ion reactivity with the coordination of the central sulfur atom, i.e., the number of S–F bonds plus two times the number of S=O bonds. Only ions with a sulfur coordination of 4 or 6 are reactive, although the reaction rate constants are generally small. The reactivity trends appear to be partially explained by spin conservation. These reactions are all sufficiently slow, so O3 reactions should not play a major role in SF6/O2 discharges. All ions studied have been found to be unreactive with O2.  相似文献   

19.
The behavior of [Fe2(CO)42‐PNPR)(μ‐pdt)] (PNPR=(Ph2PCH2)2NR, R=Me ( 1 ), Ph ( 2 ); pdt=S(CH2)3S) in the presence of acids is investigated experimentally and theoretically (using density functional theory) in order to determine the mechanisms of the proton reduction steps supported by these complexes, and to assess the role of the PNPR appended base in these processes for different redox states of the metal centers. The nature of the R substituent of the nitrogen base does not substantially affect the course of the protonation of the neutral complex by CF3SO3H or CH3SO3H; the cation with a bridging hydride ligand, 1 μH+ (R=Me) or 2 μH+ (R=Ph) is obtained rapidly. Only 1 μH+ can be protonated at the nitrogen atom of the PNP chelate by HBF4?Et2O or CF3SO3H, which results in a positive shift of the proton reduction by approximately 0.15 V. The theoretical study demonstrates that in this process, dihydrogen can be released from a η2‐H2 species in the FeIFeII state. When R=Ph, the bridging hydride cation 2 μH+ cannot be protonated at the amine function by HBF4?Et2O or CF3SO3H, and protonation at the N atom of the one‐electron reduced analogue is also less favored than that of a S atom of the partially de‐coordinated dithiolate bridge. In this situation, proton reduction occurs at the potential of the bridging hydride cation, 2 μH+ . The rate constants of the overall proton reduction processes are small for both complexes 1 and 2 (kobs≈4–7 s?1) because of the slow intramolecular proton migration and H2 release steps identified by the theoretical study.  相似文献   

20.
Preparation of Tetramethylammonium Azidosulfite and Tetramethylammonium Cyanate Sulfur Dioxide‐Adduct, [(CH3)4N]+[SO2N3], [(CH3)4N]+[SO2OCN] and Crystal Structure of [(CH3)4N]+[SO2N3] Tetramethylammonium azide forms with sulfur dioxide an azidosulfite salt. It is characterized by NMR and vibrational spectroscopy and the crystal structure analysis. [(CH3)4N]+[SO2N3] crystallizes in the monoclinic space group P21/c with a = 551.3(1) pm, b = 1095.2(1) pm, c = 1465.0(1) pm, β = 100.63(1)°, and four formula units in the unit cell. The crystal structure possesses a strong S–N interaction between the N3– anions and the SO2 molecules. The S–N distance of 200.5(2) pm is longer than a covalent single S–N bond. The structure is compared with ab initio calculated data. Furthermore an adduct of tetrametylammonium cyanate and sulfur dioxide is reported. It is characterised by NMR and vibrational spectroscopy. The structure is calculated by ab initio methods.  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