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.
Energies of homolytic cleavage of O-H bonds in 33 compounds of the general formula Ro n H (n = 2, 3, and 4) were calculated by the AMI method. For hydrotrioxides and hydrotetroxides, the bond dissociation energies are virtually independent of the nature of the substituent R:D(RO n -H) = 92.3±0.8 kcal mol–1 (n = 3 and 4).  相似文献   

2.
The thermal dehydration-decomposition of Ln2(SeO4)3·nH2O (wheren=12 forLn=Pr, Nd andn=8 forLn=Sm) and PrxLn2−x(SeO4)3·nH2O (wheren=12 forx=1.0 andLn=Nd;n=8 forx=0.2 and 1.0 in case ofLn=Sm) have been reported.
Zusammenfassung Die thermische Dehydratation-Zersetzung von Ln2(SeO4)3·nH2O (mitn=12 fürLn=Pr, Nd undn=8 fürLn=Sm) und PrxLn2−x(SeO4)3·nH2O (mitn=12 fürx=1.0 undLn=Nd;n=8 fürx=0.2 und 1.0 in Falle vonLn=Sm) wurde beschrieben.
  相似文献   

3.
本文考虑相对论效应并应用密度泛函理论(DFT)研究水溶液中UO2Xn(H2O)5-n(X=F,Cl,Br;n=1~4)和UO2Xn(X=F,Cl,Br;n=1~6)一系列水合和非水合铀酰化合物的结构和紫外吸收光谱性质。将这一系列物质命名为Xnm(X为F,Cl,和Br;n为卤素配体个数,m为水分子配体的个数)。在水溶液中,溶剂化效应采用类导体屏蔽模型(COSMO)并采用SAS溶剂接触曲面构造空穴模拟水溶剂对配合物的作用。配合物的紫外光谱性质采用考虑旋-轨耦合相对论效应的含时密度泛函(SO-TD-DFT)进行计算。U=O键随着F配体数目的增加而明显伸长,然而随Cl和Br配体数目的增加变化较小。随X配体数目的增加和水分子参与配位,铀与X的结合能逐渐减弱。配合物的紫外光谱计算表明铀酰氟的各种配合物并不出现特征吸收峰,而铀酰氯和铀酰溴的各种配合物均有特征吸收光谱。通过分子轨道分析可以很好解释光谱所体现的特征。  相似文献   

4.
The standard molar enthalpy of formation of crystalline dialkyldithiocarbamate chelates, [Pd(S2CNR2)2], with R=CH3 and i-C3H7, was determined through reaction-solution calorimetry in 1,2-dichloroethane, at 298 K. Using the standard molar enthalpies of formation of the gaseous chelates, the homolytic (526±18 and 666±10) and heterolytic (2693±18 and 2957±10 kJ mol-1) mean enthalpies of palladium-sulphur bond dissociation were calculated. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

5.
3-Picoline-N-oxide (3-PicNO) complexes of rare-earth bromides of the formulaMBr3(3-PicNO)8–n·nH2O wheren=0 forM=La, Pr, Nd, Sm Tb or Y andn=2 forM=Ho or Yb have been prepared. Infrared and proton NMR studies indicate that the coordination of the ligand is through oxygen. Conductance data in acetonitrile suggest that two bromide ions are coordinated to the metal ion. Proton NMR studies suggest a bicapped dodecahedral arrangement of the ligands around the metal ion in solution for Pr(III), Nd(III) and Tb(III) complexes.  相似文献   

6.
The electronic and spatial structures of a broad spectrum of neutral compounds with X-Hal (X = N, O, Cl; Hal = Cl, Br, I) bonds and their protonated forms and of different electronic states of triiodide cation, I3 +, were determined from density functional B3LYP/6 311G* quantum chemical calculations. The effects of the structure of these compounds on the parameters of electrophilic reactivity were revealed and the thermochemical characteristics of homolytic and heterolytic X-Hal bond dissociation and of iodine transfer in hydroxyl-containing solvents were calculated. Due to low homolytic bond dissociation energies of X-I, the formation of molecular iodine and triiodide cation I3 + becomes thermodynamically favorable and the cation should act as iodinating agent alternative to acylhypoiodites and N-iodoimides. The solvation effects of MeOH and CH2Cl2 on the X-Hal bond homolysis and heterolysis were determined using the PCM model. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1280–1288, August, 2006.  相似文献   

7.
The standard molar enthalpies of formation of crystalline dialkyldithiocarbamates chelates, [Pd(S2CNR2)2], with R=C2H5, n-C3H7, n-C4H9 and i-C4H9, were determined through reaction-solution calorimetry in acetone, at 298.15 K. From the standard molar enthalpies of formation of the gaseous chelates, the homolytic (172.4±3.8, 182.5±3.2,150.9±3.1 and 162.6±3.1 kJ mol−1) and heterolytic (745.0±3.8, 803.7±3.3,834.3±3.1 and 735.2±3.0 kJ mol−1) mean palladium-sulphur bond-dissociation enthalpies were calculated. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

8.
Hydrated aluminium cations have been investigated as a photochemical model system with up to ten water molecules by UV action spectroscopy in a Fourier transform ion cyclotron resonance (FT-ICR) mass spectrometer. Intense photodissociation was observed starting at 4.5 eV for two to eight water molecules with loss of atomic hydrogen, molecular hydrogen and water molecules. Quantum chemical calculations for n=2 reveal that solvation shifts the intense 3s–3p excitations of Al+ into the investigated photon energy range below 5.5 eV. During the photochemical relaxation, internal conversion from S1 to T2 takes place, and photochemical hydrogen formation starts on the T2 surface, which passes through a conical intersection, changing to T1. On this triplet surface, the electron that was excited to the Al 3p orbital is transferred to a coordinated water molecule, which dissociates into a hydroxide ion and a hydrogen atom. If the system remains in the triplet state, this hydrogen radical is lost directly. If the system returns to singlet multiplicity, the reaction may be reversed, with recombination with the hydroxide moiety and electron transfer back to aluminium, resulting in water evaporation. Alternatively, the hydrogen radical can attack the intact water molecule, forming molecular hydrogen and aluminium dihydroxide. Photodissociation is observed for up to n=8. Clusters with n=9 or 10 occur exclusively as HAlOH+(H2O)n-1 and are transparent in the investigated energy range. For n=4–8, a mixture of Al+(H2O)n and HAlOH+(H2O)n-1 is present in the experiment.  相似文献   

9.
Reactions of phenol and hydroxyl radical were studied under the aqueous environment to investigate the antioxidant characters of phenolic compounds. M06‐2X/6‐311 + G(d,p) calculations were carried out, where proton transfers via water molecules were examined carefully. Stepwise paths from phenol + OH + (H2O)n (n = 3, 7, and 12) to the phenoxyl radical (Ph O) via dihydroxycyclohexadienyl radicals (ipso, ortho, meta, and para OH‐adducts) were obtained. In those paths, the water dimer was computed to participate in the bond interchange along hydrogen bonds. The concerted path corresponding to the hydrogen atom transfer (HAT, apparently Ph OH + OH → Ph O + H2O) was found. In the path, the hydroxyl radical located on the ipso carbon undergoes the charge transfer to prompt the proton (not hydrogen) transfer. While the present new mechanism is similar to the sequential proton loss electron transfer (SPLET) one, the former is of the concerted character. Tautomerization reactions of ortho or para (OH)C6H5=O + (H2O)n → C6H4(OH)2(H2O)n were traced with n = 2, 3, 4, and 14. The n = 3 (and n = 14) model of ortho and para was calculated to be most likely by the strain‐less hydrogen‐bond circuit.  相似文献   

10.
Density functional theory (DFT) calculations were made on the hydrolysis of hydantoin (2,4-imidazolidinedione). In the neutral hydrolysis, reacting systems composed of hydantoin and (H2O)n with n = 1+3, 2+3, 3+3, and 4+3 were adopted. Three water molecules (“+3”) participate in the in-plane hydrogen-bond circuit, and the n–3 = 1, 2, 3 or 4 water cluster works for the out-of-plane nucleophilic attack onto the carbonyl carbon of hydantoin. Transition states (TSs) involving bond interchanges prompted by proton transfers were determined. The reaction path with n = 3+3 containing N-carbamoyl glycine, N-carboxy glycine and three tetrahedral intermediates was found to be most likely. In the acid-catalyzed hydrolysis, a reacting system composed of hydantoin and H3O+(H2O)7 was employed. Ten TSs and nine intermediates were obtained. N-carbamoyl glycine and N-carboxy glycine were confirmed to be detectable stable species. The product consists of glycine, carbonic acid (not CO2), NH4+, and (H2O)5. It has the exothermic energy, whereas the product in the neutral hydrolysis is of the endothermic one for all n values. For both neutral (n = 3+3) and acid-catalyzed hydrolyses, the rate-determining steps were calculated to be for formation of the tetrahedral intermediate, HOOC-CH2-NH-C(OH)2NH2. The pattern of proton transfers along hydrogen bonds was carefully investigated.  相似文献   

11.
The standard (p° = 0.1 MPa) molar enthalpies of combustion and sublimation of 3,4,5-trimethoxyphenol were measured, respectively, by static bomb combustion calorimetry in oxygen atmosphere and by Calvet microcalorimetry. From these measurements, the standard molar enthalpy of formation in both the crystalline and gaseous phase, at T = 298.15 K, were derived: ?(643.4 ± 1.9) kJ · mol?1 and ?(518.1 ± 3.6) kJ · mol?1, respectively. Density functional theory calculations for this compound and respective phenoxyl radical and phenoxide anion were also performed using the B3LYP functional and extended basis sets, which allowed the theoretical estimation of the gaseous phase standard molar enthalpy of formation through the use of isodesmic reactions and the calculation of the homolytic and heterolytic O–H bond dissociation energies. There is good agreement between the calculated and experimental enthalpy of formation. Substituent effects on the homolytic and heterolytic O–H bond dissociation energies have been analysed.  相似文献   

12.
Cyclic water clusters (H2O)n (n = 3–12) trapped inside organic/inorganic hosts do not correspond to the global energy minimal structures. Their closed loop connections through the H‐bonds, although weakly interacting, result in diamagnetic ring currents leading to what we term “H‐bonded aromaticity.” Such H‐bonded aromaticity in supramolecular structures generalizes the formation of such stable (H2O)n molecules confined within various host systems. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2006  相似文献   

13.
The first heterolytic and homolytic N(O)-NO(2) bond dissociation energy scales of three types Y-nitro (Y = N, O) compounds and corresponding radical anions in acetonitrile were established by using titration calorimetry combined with relevant electrochemical data through proper thermodynamic cycles.  相似文献   

14.
Interest in water adsorption on cation-substituted zeolites is due to the possibility of the M n+ (H2O) + [Si-O-Al]?1 → MOH(n ? 1)+ + Si-O(H)-Al (M = metal, n = 1–3) reaction taking place. As a result of this reaction, the cation-substituted zeolite can exhibit Brønsted acid activity. The molecular adsorption of water on Zn/ZSM-5 zeolite at room temperature and the subsequent heterolytic dissociation of adsorbed water under heating have been investigated by diffuse-reflectance IR spectroscopy. For theoretical simulation of these processes, three different fragments of the ZSM-5 lattice corresponding to possible variants of the structure of the ionic site of the substituting cation have been examined. Calculations on the molecular and dissociative adsorption of water molecules on substituting Zn2+ cations have been performed by the DFT method. Two pathways of the dissociation of adsorbed H2O molecules-endothermic and exothermic ones-have been discovered, and it has been demonstrated that the spatial separation of two lattice Al ions at the Zn2+ cation site significantly affects the adsorption energy.  相似文献   

15.
The dissociation of hydrogen bromide in a small water cluster (H2O)n (n=3–5) has been studied with quantum chemical methods. The dynamics of dissociation was followed by classical molecular dynamics, and stationary points were studied in order to compute the free energy change associated with the ionization process. The nudged elastic band method was used to map out the energy profile of the reaction paths. The results show that HBr can dissociate in the presence of just four water molecules if they are in the correct configuration.The relation of our results to recent experiments is discussed.  相似文献   

16.
Summary Cerium(III) and europium(III) are extracted at 50° C by [C8H17 As(O)OH]2(CH2) n (n=6, 8, 12) inn-octanol from aqueous solutions containing NaCl and acetic acid/sodium acetate in the pH range 6.7–8.0. The extraction coefficient, , reached a maximum of: 10.0, 13.2 forn=6; 10.7, 20.0 forn=8; 4.68, 6.03 forn=12; at an aqueous equilibrium pH of: 7.65, 7.80 forn=6; 7.89, 7.95 forn=8; 7.81, 7.82 forn=12; corresponding to the extraction efficiency of: 91%, 93% forn=6; 92%, 95% forn=8; 80%, 83% forn=12; for cerium and europium, respectively.The log was found to be second power dependent for cerium and one and one-half power dependent for europium on the negative log of the aqueous equilibrium [H+] forn=6, 8, 12. The extractions were done at 50°C using organic reagent solutions (1.09×10–3 M forn=12, 1.08 × 10–3 M forn=8, 1.32×10–3 M forn=6) and aqueous solutions of 1.0×10–5 M for CeCl3 and EuCl3, having their Cl ionic strengths adjusted to 0.100 molar with NaCl. It was found that the slopes increase in the pH dependence studies in the seriesn=6, 8, and 12 for both cerium and europium metal ions. The reagent dependence studies were done at 50°C and pH: 7.00, 6.80 forn=6; 7.30, 7.20 forn=8; 7.45, 7.35 forn=12; for cerium and europium, respectively. The log was found to be first power dependent on the log of the uncomplexed reagent concentration at equilibrium. It is assumed that the complexes are hexacoordinated.
Polymethylenebis-Oktylarsinsäuren als Reagenzien für die Flüssig-flüssig-Extraktion von Cer(III) und Europium(III)
Zusammenfassung Cer(III) und Europium(III) werden aus wäßrigen, Kochsalz und Essigsäure/Acetat enthaltenden Lösungen bei 50°C mit C8H17As(O)OH2(CH2)n (n=6, 8, 12) im pH-Gebiet 6,7–8,0 extrahiert. Der Extraktionskoeffizient erreicht ein Maximum von: 10,0, 13,2 fürn=6; 10,7, 20,0 fürn=8; 4,68, 6,03 fürn=12; bei einem wäßrigen Gleichgewicht-pH von: 7,65, 7,80 fürn=6; 7,89, 7,95 fürn=8; 7,81, 7,82 fürn=12, entsprechend dem Aus maß der Extraktion von: 91%, 93% fürn=6; 92%, 95% fürn=8; 80%, 83% fürn=12; für Ce bzw. Eu.Für Cer ist der log in zweiter Potenz, für Europium in erster und 1/2 Potenz abhängig vom negativen log des wäßrigen [H+]-Gleichgewichtes fürn=6, 8, 12. Die Extraktionen wurden bei 50°C mit organischen Reagens-lösungen (1,09×10–3 M fürn=12, 1,08×10–3 M fürn=8, 1,32×10–3 M fürn=6) und mit wäßrigen Lösungen von 1,0×10–5 M für CeCl3 und EuCl3 durchgeführt, deren lonenstärke für Cl mit NaCl auf 0,100M eingestellt wurde. Bei der Untersuchung der pH-Abhängigkeit bei den Reihenn=6, 8 und 12 ergab sich ein Anstieg sowohl für Ce wie für Eu. Die Reagens-Abhängigkeit wurde bei 50°C und bei pH 7,00, 6,80 fürn=6; 7,30, 7,20 fürn=8; 7,45, 7,35 fürn=12 für Cer bzw. Europium geprüft. Der log war einfach abhängig vom log der Konzentration des unkomplexierten Reagens beim Gleichgewicht. Es wird angenommen, daß die Komplexe hexakoordiniert sind.
  相似文献   

17.
Benzoylhydrazine, isonitrosoacetone react in ethanol with cobalt, nickel, copper halides andacetates to give the coordination compounds MX2(HL) · nH2O, MX2(HL)2 · nH2O, and ML2 · nH2O [M = Co,Ni, Cu; X = Cl, Br; HL = C6H5C(O)NHNC(CH3)CHNOH; n = 1-4]. The same reactions performed in the presence of sodium acetate and pyridine (pH 7-8) yield the complexes MLCl · nH2O and MPy2LCl ·nH2O (n = 1-4). All the complexes have a pseudooctahedral coordination geometry. The hydrazone HL behaves as a tridentate O,N,N-donor ligand. Thermolysis of the substances involves dehydration (70-90°C), deaquation (155-170°C) or deamination (175-195°C), and complete thermal decomposition (285-360°C).  相似文献   

18.
To probe the kinetic performance of microsolvated α‐nucleophile, the G2(+)M calculations were carried out for the gas‐phase SN2 reactions of monohydrated and dihydrated α‐oxy‐nucleophiles XO?(H2O)n = 1,2 (X = HO, CH3O, F, Cl, Br), and α‐sulfur‐nucleophile, HSS?(H2O)n = 1,2, toward CH3Cl. We compared the reactivities of hydrated α‐nucleophiles to those of hydrated normal nucleophiles. Our calculations show that the α‐effect of monohydrated and dihydrated α‐oxy‐nucleophiles will become weaker than those of unhydrated ones if we apply a plot of activation barrier as a function of anion basicity. Whereas the enhanced reactivity of monohydrated and dihydrated ROO? (R = H, Me) could be observed if compared them with the specific normal nucleophiles, RO? (R = H, Me). This phenomena can not be seen in the comparisons of XO?(H2O)n = 1,2 (X = F, Cl, Br) with ClC2H4O?(H2O)n = 1,2, a normal nucleophile with similar gas basicity to XO?(H2O)n = 1,2. These results have been carefully analyzed by natural bond orbital theory and activation strain model. Meanwhile, the relationships between activation barriers with reaction energies and the ionization energies of α‐nucleophile are also discussed. © 2015 Wiley Periodicals, Inc.  相似文献   

19.
Here we report the synthesis, structure and detailed characterisation of three n‐membered oxovanadium rings, Nan[(V=O)nNan(H2O)n(α, β, or γ‐CD)2]?m H2O (n=6, 7, or 8), prepared by the reactions of (V=O)SO4?x H2O with α, β, or γ‐cyclodextrins (CDs) and NaOH in water. Their alternating heterometallic vanadium/sodium cyclic core structures were sandwiched between two CD moieties such that O‐Na‐O groups separated the neighbouring vanadyl ions. Antiferromagnetic interactions between the S=1/2 vanadyl ions led to S=0 ground states for the even‐membered rings, but to two quasi‐degenerate S=1/2 states for the spin‐frustrated heptanuclear cluster.  相似文献   

20.
The stability, infrared spectra and electronic structures of (ZrO2)n (n=3–6) clusters have been investigated by using density‐functional theory (DFT) at B3LYP/6‐31G* level. The lowest‐energy structures have been recognized by considering a number of structural isomers for each cluster size. It is found that the lowest‐energy (ZrO2)5 cluster is the most stable among the (ZrO2)n (n=3–6) clusters. The vibration spectra of Zr? O stretching motion from terminal oxygen atom locate between 900 and 1000 cm?1, and the vibrational band of Zr? O? Zr? O four member ring is obtained at 600–700 cm?1, which are in good agreement with the experimental results. Mulliken populations and NBO charges of (ZrO2)n clusters indicate that the charge transfers occur between 4d orbital of Zr atoms and 2p orbital of O atoms. HOMO‐LUMO gaps illustrate that chemical stabilities of the lowest‐energy (ZrO2)n (n=3–6) clusters display an even‐odd alternating pattern with increasing cluster size.  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