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 anaerobic treatment of raw vinasse in a combined system consisting in two methanogenic reactors, up-flow anaerobic sludge blanket (UASB) + anaerobic packed bed reactors (APBR), was evaluated. The organic loading rate (OLR) was varied, and the best condition for the combined system was 12.5 kg COD m?3day?1 with averages of 0.289 m3 CH4 kg COD r?1for the UASB reactor and 4.4 kg COD m?3day?1 with 0.207 m3 CH4 kg COD r?1 for APBR. The OLR played a major role in the emission of H2S conducting to relatively stable quality of biogas emitted from the APBR, with H2S concentrations <10 mg L?1. The importance of the sulphate to COD ratio was demonstrated as a result of the low biogas quality recorded at the lowest ratio. It was possible to develop a proper anaerobic digestion of raw vinasse through the combined system with COD removal efficiency of 86.7% and higher CH4 and a lower H2S content in biogas.  相似文献   

2.
The [C4H6O] ion of structure [CH2?CHCH?CHOH] (a) is generated by loss of C4H8 from ionized 6,6-dimethyl-2-cyclohexen-1-ol. The heat of formation ΔHf of [CH2?CHCH?CHOH] was estimated to be 736 kJ mol?1. The isomeric ion [CH2?C(OH)CH?CH2] (b) was shown to have ΔHf, ? 761 kJ mol?1, 54 kJ mol?1 less than that of its keto analogue [CH3COCH?CH2]. Ion [CH2?C(OH)CH?CH2] may be generated by loss of C2H4 from ionized hex-1-en-3-one or by loss of C4H8 from ionized 4,4-dimethyl-2-cyclohexen-1-ol. The [C4H6O] ion generated by loss of C2H4 from ionized 2-cyclohexen-1-ol was shown to consist of a mixture of the above enol ions by comparing the metastable ion and collisional activation mass spectra of [CH2?CHCH?CHOH] and [CH2?C(OH)CH?CH2] ions with that of the above daughter ion. It is further concluded that prior to their major fragmentations by loss of CH3˙ and CO, [CH2?CHCH?CHOH]+˙ and [CH2?C(OH)CH?CH2] do not rearrange to their keto counterparts. The metastable ion and collisional activation characteristics of the isomeric allenic [C4H6O] ion [CH2?C?CHCH2OH] are also reported.  相似文献   

3.
A three‐dimensional (3D) cage‐like organic network (3D‐CON) structure synthesized by the straightforward condensation of building blocks designed with gas adsorption properties is presented. The 3D‐CON can be prepared using an easy but powerful route, which is essential for commercial scale‐up. The resulting fused aromatic 3D‐CON exhibited a high Brunauer–Emmett–Teller (BET) specific surface area of up to 2247 m2 g?1. More importantly, the 3D‐CON displayed outstanding low pressure hydrogen (H2, 2.64 wt %, 1.0 bar and 77 K), methane (CH4, 2.4 wt %, 1.0 bar and 273 K), and carbon dioxide (CO2, 26.7 wt %, 1.0 bar and 273 K) uptake with a high isosteric heat of adsorption (H2, 8.10 kJ mol?1; CH4, 18.72 kJ mol?1; CO2, 31.87 kJ mol?1). These values are among the best reported for organic networks with high thermal stability (ca. 600 °C).  相似文献   

4.
The mechanisms for the reaction of CH3SSCH3 with OH radical are investigated at the QCISD(T)/6‐311++G(d,p)//B3LYP/6‐311++G(d,p) level of theory. Five channels have been obtained and six transition state structures have been located for the title reaction. The initial association between CH3SSCH3 and OH, which forms two low‐energy adducts named as CH3S(OH)SCH3 (IM1 and IM2), is confirmed to be a barrierless process, The S? S bond rupture and H? S bond formation of IM1 lead to the products P1(CH3SH + CH3SO) with a barrier height of 40.00 kJ mol?1. The reaction energy of Path 1 is ?74.04 kJ mol?1. P1 is the most abundant in view of both thermodynamics and dynamics. In addition, IMs can lead to the products P2 (CH3S + CH3SOH), P3 (H2O + CH2S + CH3S), P4 (CH3 + CH3SSOH), and P5 (CH4 + CH3SSO) by addition‐elimination or hydrogen abstraction mechanism. All products are thermodynamically favorable except for P4 (CH3 + CH3SSOH). The reaction energies of Path 2, Path 3, Path 4, and Path 5 are ?28.42, ?46.90, 28.03, and ?89.47 kJ mol?1, respectively. Path 5 is the least favorable channel despite its largest exothermicity (?89.47 kJ mol?1) because this process must undergo two barriers of TS5 (109.0 kJ mol?1) and TS6 (25.49 kJ mol?1). Hopefully, the results presented in this study may provide helpful information on deep insight into the reaction mechanism. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

5.
Ab initio molecular orbital calculations with split-valence plus polarization basis sets and incorporating valence-electron correlation have been performed to determine the equilibrium structure of ethyloxonium ([CH3CH2OH2]+) and examine its modes of unimolecular dissociation. An asymmetric structure (1) is predicted to be the most stable form of ethyloxonium, but a second conformational isomer of Cs symmetry lies only 1.4 kJ mol?1 higher in energy than 1. Four unimolecular decomposition pathways for 1 have been examined involving loss of H2, CH4, H2O or C2H4. The most stable fragmentation products, lying 65 kJ mol?1 above 1, are associated with the H2 elimination reaction. However, large barriers of 257 and 223 kJ mol?1 have to be surmounted for H2 and CH4 loss, respectively. On the other hand, elimination of either C2H4 or H2O from ethyloxonium can proceed without a barrier to the reverse associations and, with total endothermicities of 130 and 160 kJ mol?1, respectively, these reactions are expected to dominate at lower energies. A second important equilibrium structure on the surface is a hydrogen-bridged complex, lying 53 kJ mol?1 above 1. This complex is involved in the C2H4 elimination reaction, acts as an intermediate in the proton-transfer reaction connecting [C2H5]+ +H2O and C2H4 + [H3O]+ and plays an important role in the isotopic scrambling that has been observed experimentally in the elimination of either H2O or C2H4 from ethyloxonium. The proton affinity of ethanol was calculated as 799 kJ mol?1, in close agreement with the experimental value of 794 kJ mol?1.  相似文献   

6.
Multi-phase anaerobic reactor for H2 and CH4 production from paperboard mill wastewater was studied. The reactor was operated at hydraulic retention times (HRTs) of 12, 18, 24, and 36 h, and organic loading rates (OLRs) of 2.2, 1.5, 1.1, and 0.75 kg chemical oxygen demand (COD)/m3 day, respectively. HRT of 12 h and OLR of 2.2 kg COD/m3 day provided maximum hydrogen yield of 42.76?±?14.5 ml/g CODremoved and volumetric substrate uptake rate (?rS) of 16.51?±?4.43 mg COD/L h. This corresponded to the highest soluble COD/total COD (SCOD/TCOD) ratio of 56.25?±?3.3 % and the maximum volatile fatty acid (VFA) yield (YVFA) of 0.21?±?0.03 g VFA/g COD, confirming that H2 was mainly produced through SCOD conversion. The highest methane yield (18.78?±?3.8 ml/g CODremoved) and ?rS of 21.74?±?1.34 mgCOD/L h were achieved at an HRT of 36 h and OLR of 0.75 kg COD/m3 day. The maximum hydrogen production rate (HPR) and methane production rate (MPR) were achieved at carbon to nitrogen (C/N) ratio of 47.9 and 14.3, respectively. This implies the important effect of C/N ratio on the distinction between the dominant microorganism bioactivities responsible for H2 and CH4 production.  相似文献   

7.
The vulcanization of rubber by sulfur is a large‐scale industrial process that is only poorly understood, especially the role of zinc oxide, which is added as an activator. We used the highly symmetrical cluster Zn4O4 (Td) as a model species to study the thermodynamics of the initial interaction of various vulcanization‐related molecules with ZnO by DFT methods, mostly at the B3LYP/6‐31+G* level. The interaction energy of Lewis bases with Zn4O4 increases in the following order: CO62H43H62S2<1,4‐C5H82O2S3N?CH3COO?. The corresponding binding energies range from ?57 to ?262 kJ mol?1. However, Brønsted acids react with the Zn4O4 cluster with proton transfer from the ligand molecule to one of the oxygen atoms of Zn4O4, and these reactions are all strongly exothermic [binding energies [kJ mol?1] in parentheses: H2O (?183), MeOH (?171), H2S (?245), MeSH (?230), C3H6 (?121), and CH3COOH (?255)]. The important vulcanization accelerator mercaptobenzothiazole (C7H5NS2, MBT) containing several donor sites reacts with the Zn4O4 cluster with proton transfer from the NH group to one of the oxygen atoms of ZnO, and in addition the exocyclic thiono sulfur atom and the nitrogen atom coordinate to one and the same zinc atom, resulting in a binding energy of ?247 kJ mol?1. A second isomer of [(MBT)Zn4O4] with a strong O? H???N hydrogen bond rather than a Zn? N bond is only slightly less stable (binding energy ?243 kJ mol?1). The NH form of free MBT is 36 kJ mol?1 more stable than the tautomeric SH form, while the sulfurized MBT derivative benzothiazolyl hydrodisulfide C7H5NS3 (BtSSH) is most stable with the connectivity >CSSH.  相似文献   

8.
Ab initio molecular orbital calculations with moderately large polarization basis sets and including valence-electron correlation have been used to examine the structure and dissociation mechanisms of protonated methanol [CH3OH2]+. Stable isomers and transition structures have been characterized using gradient techniques. Protonated methanol is found to be the only stable isomer in the [CH5O]+ potential surface. There is no evidence for a tightly-bound complex, [HOCH2]+…?H2, analogous to the preferred structure [CH3]+…?H2 of [CH5]+. Protonated methanol is found to possess a pyramidal arrangement of bonds at the oxygen atom with a barrier to inversion of 8kJ mol?1. The lowest energy fragmentation pathways are dissociation into methyl cation and water (predicted to require 284 kJ mol?1 with zero reverse activation energy) and loss of molecular hydrogen (endothermic by 138 kJ mol?1 but with a reverse activation barrier of 149 kJ mol?1). The results offer a possible explanation as to why production of [CH2OH]+ from the reaction of methyl cation with water is not observed. Other dissociation processes examined include loss of a hydrogen atom to yield the methylenoxonium radical cation or methanol radical cation (requiring 441 and 490 kJ mol?1, respectively) and loss of a proton to yield neutral methanol (requiring 784 kJ mol?1).  相似文献   

9.
The following isomers of the ethyl halide molecular ions have all been shown to be stable species in the gas phase: [CH2CH2FH]+˙; [CH3ClCH2]+˙ (ΔHf° = 1012 kJ mol?1); [CH3CHClH]+˙ (ΔHf° = 971 kJ mol?1); [CH2CH2ClH]+˙; [CH3BrCH2]+˙ (ΔHf° = 1058 KJ mol?1); [CH3CHBrH]+˙ (ΔHf° = 995 kJ mol?1) and [CH2CH2BrH]+˙. Neutralization–reionization mass spectrometry, employing Xe as the electron transfer target gas and O2 as the target gas for reionization, indicated that the ylides CH3ClCH2 and CH3BrCH2 could not be generated by such means. However, the species CH3CHClH, CH2CH2ClH and CH2CH2BrH (and possibly CH3CHBrH) were unambiguously identified.  相似文献   

10.
The performance and microbial community characteristics of a laboratory scale anaerobic baffled reactor (ABR) with four compartments (C1–C4) treating sugar refinery wastewater were investigated. The COD removal was 94.8 % with a CH4 yield of 0.21 L?g?1 CODremoved at total organic loading rate (OLR) of 5.33 kg COD/m?3?day?1. Fermentative bacteria were dominant in C1 and C2, while syntrophic acetogens and methanogens were dominant in C3 and C4. Some acid-tolerant methanogens were enriched in acidogenic phase. The present of the acid-tolerant methanogens could improve the efficiency and stability of the ABR as the most of the methanogens are vulnerable to low pH. In addition, high functional redundancy of the fermentative bacteria implicated that the microbial communities in acidogenic phase were stable functionally and allowed the ABR to balance perturbation. It was also found that syntrophic acetogenesis might be a weakness in the ABR as syntrophic acetogens were poor as compared with fermentative bacteria and methanogens.  相似文献   

11.
[RuCl2(NCCH3)2(cod)], an alternative starting material to [RuCl2(cod)] n for the preparation of ruthenium(II) complexes, has been prepared from the polymer compound and isolated in yields up to 87% using a new work-up procedure. The compound has been obtained as a yellow solid without water of crystallization. The complexes [RuCl2(NCR)2(cod)] spontaneously transform into dimers [Ru2Cl(μ-Cl)3(cod)2(NCR)] (R?=?Me, Ph). 1H NMR kinetic experiments for these transformations evidenced first-order behavior. [RuCl2(NCPh)2(cod)] dimerizes slower by a factor of ten than [RuCl2(NCCH3)2(cod)]. The following activation parameters, ΔH #?=?114?±?3?kJ?mol?1 and ΔS #?=?66?±?9?J?K?1?mol?1 for R?=?CH3CN (ΔG #?=?94?±?5?kJ?mol?1, 298.15?K) and ΔH #?=?122?±?2?kJ?mol?1 and ΔS #?=?75?±?6?J?K?1?mol?1 for R?=?Ph (ΔG #?=?100?±?4?kJ?mol?1, 298.15?K), have been calculated from the first-order rate constants in the temperature range 294–323?K. The kinetic parameters are in agreement with a two-step mechanism with dissociation of acetonitrile as the rate-determining step. The molecular structures of [Ru2Cl(μ-Cl)3(cod)2(NCR)] (R?=?Me, Ph) have been determined by X-ray diffraction.  相似文献   

12.
Polyhydroxy polyurethane sorbent was modified by the addition of halogen atoms to its matrix to produce a new sorbent distinguished by high surface polarity, enhanced capacity, and improved stability in both acidic and alkaline media. Halo polyhydroxy polyurethane foam (X-PPF) was characterized by NMR, FTIR, UV–Vis, Raman spectroscopy, pHZCP values, and scanning electron microscopy images. Experimental studies have proven that X-PPFs have a great potential for the extraction and recovery of cobalt ions and this was attributed to the presence of halogen, phenolic, and urethane groups. The pHZCP value of X-PPFs was determined to be 0.91 and the maximum metal recovery was achieved at a pH range of 6–7. The kinetics of the process was best described by pseudo-second-order model (R2?=?1). ΔH, ΔS, and ΔG values were calculated to be ?57.2?kJ?mol?1, ?172.6?J?K?1?mol?1, and ?5.8?kJ?mol?1, respectively. A perfect isotherm curve with zero intercept (0.002), good correlation (R2?=?0.999), and capacity of 246.8?mg?g?1 was obtained.  相似文献   

13.
Ab initio molecular orbital calculations have been used to study the condensation reactions of CH3? with NH3, H2O, HF and H2S. Geometry optimization has been carried out at the Hartree—Fock (HF) level with the split-valence plus d-polarization 6-31G* basis set and improved relative energies obtained from calculations which employ the split-valence plus dp-polarization 6-31G** basis set with electron correlation incorporated via Moller—Plesset perturbation theory terminated at third order (MP3). Zero-point vibrational energies have also been determined and taken into account in deriving relative energies. The structures of the intermediates CH3XH? (X = NH2, OH, F and SH) have been obtained and dissociation of these intermediates into CH2X+ + H2 on the one hand, and CH3? + HX on the other, has been examined. It is found that for those species for which the methyl condensation reaction is observed to have an appreciable rate (X = NH2 and SH), the transition structure for hydrogen elimination from CH3XH? lies significantly lower in energy than the reactants CH3? + HX (by 75 and 70 kJ mol?1 respectively). On the other hand, for those species for which the methyl condensation reaction is not observed (X = OH and F), the transition structure for H2 elimination lies higher in energy than CH3? + HX (by 6 and 87 kJ mol?1 respectively).  相似文献   

14.
Critical energies for 1,3-R sigmatropic migrations (R?H, OH, CH3 and C6H5) have been calculated by means of the MINDO/3 method. This investigation was carried out for the system [RCH2CH?CH2] and the results indicated that hydroxyl group migration is energetically favoured (critical energy = 53 kJ mol?1). Calculations are also presented for [3-buten-2-ol] isomerizations. The lowest energy pathway is related to the [2-buten-1-ol]; the corresponding critical energy for OH migration is equal to 33 kJ mol?1 in this case.  相似文献   

15.
A dynamic column breakthrough (DCB) apparatus was used to study the separation of CH4+N2 gas mixtures using two zeolites, H+-mordenite and 13X, at temperatures of (229.2 and 301.9)?K and pressures to 792.9?kPa. The apparatus is not limited to the study of dilute adsorbates within inert carrier gases because the instrumentation allows the effluent flow rate to be measured accurately: a method for correcting apparent effluent mass flow readings for large changes in effluent composition is described. The mathematical framework used to determine equilibrium adsorption capacities from the dynamic adsorption experiments is presented and includes a method for estimating quantitatively the uncertainties of the measured capacities. Dynamic adsorption experiments were conducted with pure CH4, pure N2 and equimolar CH4+N2 mixtures, and the results were compared with similar static adsorption experiments reported in the literature. The 13X zeolite had the greater adsorption capacity for both CH4 and N2. At 792?kPa the equilibrium capacities of the 13X zeolite increased from 2.13±0.14?mmol?g?1 for CH4 and 1.36±0.10?mmol?g?1 for N2 at 301.9?K to 3.97±0.19?mmol?g?1 for CH4 and 3.33±0.12?mmol?g?1 for N2 at 229.2?K. Both zeolites preferentially adsorbed CH4; however, the mordenite had a greater equilibrium selectivity of 3.5±0.4 at 301.9?K. Equilibrium selectivities inferred from pure fluid capacities using the Ideal Adsorbed Solution theory were limited by the accuracy of the literature pure fluid Toth models. Equilibrium capacities with quantitative uncertainties derived directly from DCB measurements without reference to a dynamic model should help increase the accuracy of mass transfer parameters extracted by the regression of such models to time dependent data.  相似文献   

16.
The complex formation reactions of [Cu(NTP)(OH2)]4? (NTP?=?nitrilo-tris(methyl phosphonic acid)) with some selected bio-relevant ligands containing different functional groups, are investigated. Stoichiometry and stability constants for the complexes formed are reported. The results show that the ternary complexes are formed in a stepwise mechanism whereby NTP binds to copper(II), followed by coordination of amino acid, peptide or DNA. Copper(II) is found to form Cu(NTP)H n species with n?=?0, 1, 2 or 3. The concentration distribution of the various complex species has been evaluated. The kinetics of base hydrolysis of glycine methyl ester in the presence of copper(II)-NTP complex is studied in aqueous solution at different temperatures. It is proposed that the catalysis of GlyOMe ester occurs by attack of OH? ion on the uncoordinated carbonyl carbon atom of the ester group. Activation parameters for the base hydrolysis of the complex [Cu(NTP)NH2CH2CO2Me]4? are, ΔH±?=?9.5?±?0.3?kJ?mol?1 and ΔS±?=??179.3?±?0.9?J?K?1?mol?1. These show that catalysis is due to a substantial lowering of ΔH±.  相似文献   

17.
Two new doubly methoxido-bridged MnIII dinuclear complexes, [MnIII(mphp)(μ-OCH3)(CH3OH)]2·2CH3OH (1) and ([MnIII(ahbz)(μ-OCH3)(CH3OH)]2·2CH3OH (2), have been synthesized by using the tridentate ligands H2mphp (H2mphp = 2-methyl-6-(pyrimidin-2-yl-hydrazonomethyl)-phenol) and H2ahbz (H2ahbz = N-(2-amino-propyl)-2-hydroxy-benzamide). The complexes have been characterized by single-crystal X-ray diffraction analysis and magnetic measurements. Complexes 1 and 2 have a similar dimeric molecular structure. Two [Mn(L)(CH3OH)]+ moieties (L2? = mphp2? or ahbz2?) are bridged by two μ-OCH3? groups in the axial-equatorial asymmetric manner. The coordination geometry of MnIII is an axially elongated octahedron with two oxygens of a methanol ligand and a methoxido ligand situated at the axial positions. Magnetic measurements indicate that 1 and 2 exhibit antiferromagnetic behavior with the fitting parameter of J = ?1.49(3) cm?1, D = ?1.3(1) cm?1, g = 1.98(1) and zJ′ = ?0.18(4) cm?1 for 1, and J = ?1.6(2) cm?1, D = 4.5(3) cm?1, g = 2.06(1) and zJ′ = 1.4(1) cm?1 for 2 on the basis of the spin Hamiltonian ? = ?2J?Mn1?Mn2.  相似文献   

18.
IR photodissociation spectra of mass‐selected clusters composed of protonated benzene (C6H7+) and several ligands L are analyzed in the range of the C? H stretch fundamentals. The investigated systems include C6H7+? Ar, C6H7+? (N2)n (n=1–4), C6H7+? (CH4)n (n=1–4), and C6H7+? H2O. The complexes are produced in a supersonic plasma expansion using chemical ionization. The IR spectra display absorptions near 2800 and 3100 cm?1, which are attributed to the aliphatic and aromatic C? H stretch vibrations, respectively, of the benzenium ion, that is, the σ complex of C6H7+. The C6H7+? (CH4)n clusters show additional C? H stretch bands of the CH4 ligands. Both the frequencies and the relative intensities of the C6H7+ absorptions are nearly independent of the choice and number of ligands, suggesting that the benzenium ion in the detected C6H7+? Ln clusters is only weakly perturbed by the microsolvation process. Analysis of photofragmentation branching ratios yield estimated ligand binding energies of the order of 800 and 950 cm?1 (≈9.5 and 11.5 kJ mol?1) for N2 and CH4, respectively. The interpretation of the experimental data is supported by ab initio calculations for C6H7+? Ar and C6H7+? N2 at the MP 2/6‐311 G(2df,2pd) level. Both the calculations and the spectra are consistent with weak intermolecular π bonds of Ar and N2 to the C6H7+ ring. The astrophysical implications of the deduced IR spectrum of C6H7+ are briefly discussed.  相似文献   

19.
Enthalpy, activation energy, and rate constant of 9 alkyl, 3 acyl, 3 alkoxyl, and 9 peroxyl radicals with alkanethiols, benzenethiol, and L ‐cysteine are calculated. The intersection parabolas model is used for activation energy calculations. Depending on the structure of attacking radical, the activation energy of reactions with alkylthiols varies from 3 to 43 kJ mol?1 for alkyl radicals, from 7 to 9 kJ mol?1 for alkoxyl, and from 18 to 35 kJ mol?1 for peroxyl radicals. The influence of adjacent π‐bonds on activation energy is estimated. The polar effect is found in reactions of hydroxyalkyl and acyl radicals with alkylthiols. The steric effect is observed in reactions of alkyl radicals with tert‐alkylthiols. All these factors are characterized via increments of activation energy. Quantum chemical calculations of activation energy and geometry of transition state were performed for model reactions: C?H3 + CH3SH, CH3O? + CH3SH, and HO2? + CH3SH with using density functional theory and Gaussian‐98. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 41: 284–293, 2009  相似文献   

20.
(1,5-Cyclooctadiene) (4-substituted pyridinium 2-pyridylcarbonylmethylide)- rhodium(I) perchlorates, [Rh(COD)(C5H4NC(O)C?H+C5H4X-4)]ClO4 [COD = 1,5-cyclooctadiene; X = CH3C(O), CH3OC(O), C6H5, CH3, and H], have been prepared. They are shown to have the geometry with coordination by the pyridyl nitrogen and carbonyl oxygen atoms of the ylide ligands and to exhibit intramolecular rearrangement of coordinated COD in chloroform, methanol, and dimethyl sulphoxide based on IR and 1H NMR spectroscopies. Although the ylides have exhibited fluorescence bands due to an intramolecular charge-transfer transition and phosphorescence bands due to a carbonyl 3(n*) transition, the complexes have given emission bands due to the metal-to-ylide ligand charge-transfer transition. A.single crystal X-ray crystal structure has been determined for [Rh(COD)(C5H4NC(O)C?H+C5H4CH3-4)]ClO4. The crystals are monoclinic, space group P21/n with cell dimensions a = 14.887(3), b = 20.274(4), c = 6.966(1) Å, β = 96.13(1)°, and Z = 4. The structure has been refined by a block-diagonal least-squares method to final R = 0.060 for 2997 independent reflections with |Fo| > 3σ(F). The ylide carbon-pyridinium nitrogen bond distance is 1.420(10) Å. The bonded distances from rhodium to the midpoints of the double bonds of COD are 1.982(11) and 2.014(12) Å.  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