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.
Here a unique single-crystal-to-single-crystal (SCSC) transformation of a 116-nuclear AuI72CdII40NaI4 cage-of-cage ( 2 CdNa) is reported, which was created from a trigold(I) metalloligand with d -penicillamine by way of a 9-nuclear AuI6CdII3 cage ( 1 ). Cage-of-cage 2 CdNa is composed of 12 cages of 1 that are linked by 4 Cd2+ and 4 Na+ ions, with its surface being covered by 12 NO3 ions to form a discrete, spherical molecule with a diameter ca. 4.7 nm. In crystal 2 CdNa, the cage-of-cage molecules are packed in a cubic lattice with a huge cell volume of ca. 4.5×105 Å3, so as to have large interstices with diameters of more than 3 nm. Upon soaking crystals 2 CdNa in aqueous Cu(NO3)2, all Cd2+ and Na+ were quickly exchanged by Cu2+ to produce an analogous AuI72CuII44 cage-of-cage ( 2 Cu) in a SCSC manner. Prolonged soaking led to the SCSC transformation to another supramolecular structure ( 2′ Cu) consisting of 152-nuclear AuI72CuII80 cage-of-cages that are alternately H-bonded with the AuI72CuII44 cage-of-cages. 2′ Cu showed the accommodation of MoO42− and the conversion of MoO42− to β-Mo8O264− in the crystal, with retention of single-crystallinity.  相似文献   

2.
为探讨急性心肌梗死(AMI)患者血清中K+、Na+、Ca2+、Fe2+、Mg2+含量变化,并研究其与心肌梗死患者之间的关系。选取2022年5月至2023年2月收治的AMI患者37例,同时选取健康体检者35例作为对照组。依据入院时或体检时收集的抽血样本进行临床生化分析,比较两组间血清K+、Na+、Ca2+、Fe2+、Mg2+含量,采用判别方程、主成分分析法(PCA),判断分析哪种金属离子对于心肌梗死的诊断价值大。结果表明,AMI患者的血清中Ca2+和Fe2+含量低于健康对照组,差异具有统计学意义。基于血钙、铁水平两组具有显著性差异,以它们为基础进行判别分析,获得判别函数式。将血清中K+、Na+、Ca2+、Fe2+、Mg2+  相似文献   

3.
Unprecedented functionalized products with an η4‐P5 ring are obtained by the reaction of [Cp*Fe(η5‐P5)] ( 1 ; Cp*=η5‐C5Me5) with different nucleophiles. With LiCH2SiMe3 and LiNMe2, the monoanionic products [Cp*Fe(η4‐P5CH2SiMe3)]? and [Cp*Fe(η4‐P5NMe2)]?, respectively, are formed. The reaction of 1 with NaNH2 leads to the formation of the trianionic compound [{Cp*Fe(η4‐P5)}2N]3?, whereas the reaction with LiPH2 yields [Cp*Fe(η4‐P5PH2)]? as the main product, with {[Cp*Fe(η4‐P5)]2PH}2? as a byproduct. The calculated energy profile of the reactions provides a rationale for the formation of the different products.  相似文献   

4.
A reliable synthesis of unstable and highly reactive BrO2F is reported. This compound can be converted into BrO2+SbF6?, BrO2+AsF6?, and BrO2+AsF6??2 BrO2F. The latter decomposes into mixed‐valent Br3O4?Br2+AsF6? with five‐, three‐, one‐, and zero‐valent bromine. BrO2+ H(SO3CF3)2? is formed with HSO3CF3. Excess BrO2F yields mixed‐valent Br3O6+OSO3CF3? with five‐ and three‐valent bromine. Reactions of BrO2F and MoF5 in SO2ClF or CH2ClF result in Cl2BrO6+Mo3O3F13?. The reaction of BrO2F with (CF3CO)2O and NO2 produces O2Br‐O‐CO‐CF3 and the known NO2+Br(ONO2)2?. All of these compounds are thermodynamically unstable.  相似文献   

5.
Synthesis, Structure, and Reactivity of Bis(dialkylamino)diphosphines Starting with the aminochlorophosphines iPr2N? PCl2 1 and (iPr2N)2P? Cl 2 , the synthesis of some new functionalized aminophosphines (iPr2N)2P? SiMe3 3a , (iPr2N)2P? SnMe3 3b , (iPr2N)(DMP)P? Cl 4 , iPr2N? P(SiMe3)2 5 and iPr2N? P(SiMe3)Cl 6 is reported. Reactions of 2 with different phosphides yield the aminodiphosphines (iPr2N)2P? P(SiMe3)2 7a , (iPr2N)2P? P(SiMe2tBu)2 7b , (iPr2N)2P? PPh2 8 and (iPr2N)2P? PH2 9 . The phosphines 3a/b react with halogenophosphines to the aminohalogenodiphosphines (iPr2N)2P? PCl2 10 , (iPr2N)2P? PtBuCl 11 and (iPr2N)2P? P(NiPr2)Cl 12 . The ambivalente aminophosphine 6 gives the aminotrichlorodiphosphine Cl(iPr2N)P? PCl2 13 after condensation with PCl3, while the reactions with the corresponding lithiumphosphides yield the aminosilyldiphosphines (iPr2N)(SiMe3)P? P(SiMe3)2 14a and (iPr2N)(SiMe3)P? P(SiMe2tBu)2 14b . The aminochlorophosphines 2/4 are reductively coupled with magnesium leading to the symmetrically substituted tetraaminodiphosphines (iPr2N)2P? P(iPr2N)2 15a and DMP(iPr2N)P? P(iPr2N)DMP 15b . The functionalized aminosilyldiphosphine 7a is treated with methanol to yield the diphosphine (iPr2N)2P? PH(SiMe3) 16 and gives the lithium phosphinophosphide (iPr2N)2P? PLi(SiMe3) 17 after metallation with n-BuLi. The compounds are characterized by their NMR and mass spectra and the 31P-NMR values of the diphosphines are discussed according to their substituents. The crystal structures of 7b, 8 and 15b showing significantly differing conformations are presented.  相似文献   

6.
The rate constants and product ion branching ratios were measured for the reactions of various small negative ions with O2(X 3Σg) and O2(a 1Δg) in a selected ion flow tube (SIFT). Only NH2 and CH3O were found to react with O2(X) and both reactions were slow. CH3O reacted by hydride transfer, both with and without electron detachment. NH2 formed both OH, as observed previously, and O2, the latter via endothermic charge transfer. A temperature study revealed a negative temperature dependence for the former channel and Arrhenius behavior for the endothermic channel, resulting in an overall rate constant with a minimum at 500 K. SF6, SF4, SO3 and CO3 were found to react with O2(a 1Δg) with rate constants less than 10−11 cm3 s−1. NH2 reacted rapidly with O2(a 1Δg) by charge transfer. The reactions of HO2 and SO2 proceeded moderately with competition between Penning detachment and charge transfer. SO2 produced a SO4 cluster product in 2% of reactions and HO2 produced O3 in 13% of the reactions. CH3O proceeded essentially at the collision rate by hydride transfer, again both with and without electron detachment. These results show that charge transfer to O2(a 1Δg) occurs readily if the there are no restrictions on the ion beyond the reaction thermodynamics. The SO2 and HO2 reactions with O2(a) are the only known reactions involving Penning detachment besides the reaction with O2 studied previously [R.S. Berry, Phys. Chem. Chem. Phys., 7 (2005) 289–290].  相似文献   

7.
Complex [(DIPePBDI)Ca]2(C6H6), with a C6H62− dianion bridging two Ca2+ ions, reacts with benzene to yield [(DIPePBDI)Ca]2(biphenyl) with a bridging biphenyl2− dianion (DIPePBDI=HC[C(Me)N-DIPeP]2; DIPeP=2,6-CH(Et)2-phenyl). The biphenyl complex was also prepared by reacting [(DIPePBDI)Ca]2(C6H6) with biphenyl or by reduction of [(DIPePBDI)CaI]2 with KC8 in presence of biphenyl. Benzene-benzene coupling was also observed when the deep purple product of ball-milling [(DIPPBDI)CaI(THF)]2 with K/KI was extracted with benzene (DIPP=2,6-CH(Me)2-phenyl) giving crystalline [(DIPPBDI)Ca(THF)]2(biphenyl) (52 % yield). Reduction of [(DIPePBDI)SrI]2 with KC8 gave highly labile [(DIPePBDI)Sr]2(C6H6) as a black powder (61 % yield) which reacts rapidly and selectively with benzene to [(DIPePBDI)Sr]2(biphenyl). DFT calculations show that the most likely route for biphenyl formation is a pathway in which the C6H62− dianion attacks neutral benzene. This is facilitated by metal-benzene coordination.  相似文献   

8.
The complexes [Cu2Br4]2?, [Cu2I4]2?, [Cu2I2Br2]2?, [Cu2I3Cl]2?, [Ag2Cl4]2? have been characterized as their isomorphous bis(triphenylphosphoranylidene)ammonium ([Ph3PNPPh3]+ = PNP+) salts by single crystal structural determinations. All anions show the centrosymmetric doubly halogen‐bridged forms [XM(μ‐X)2MX]2? with three‐coordinate metal atoms that have been observed in [M2X4]2? complexes with other large organic cations. In [Cu2I2Br2]2? the iodide ligands occupy the bridging positions and the bromide the terminal positions, while in [Cu2I3Cl]2?, obtained in an attempt to prepare [Cu2I2Cl2]2?, two of the iodide ligands occupy the bridging positions with the third iodide and the chloride ligand occupying two statistically disordered terminal positions. In [Ag2Cl4]2? the distortion from ideal trigonal coordination of the metal atom is greater than in the copper complexes, but less than in other previously reported [Ag2Cl4]2? complexes with organic cations. The ν(MX) bands have been assigned in the far‐IR spectra, and confirm previous observations regarding the unexpectedly simple IR spectra of [Cu2X4]2? complexes.  相似文献   

9.
Coordination Chemistry of P‐rich Phosphanes and Silylphosphanes XXI The Influence of the PR3 Ligands on Formation and Properties of the Phosphinophosphinidene Complexes [{η2tBu2P–P}Pt(PR3)2] and [{η2tBu2P1–P2}Pt(P3R3)(P4R′3)] (R3P)2PtCl2 and C2H4 yield the compounds [{η2‐C2H4}Pt(PR3)2] (PR3 = PMe3, PEt3, PPhEt2, PPh2Et, PPh2Me, PPh2iPr, PPh2tBu and P(p‐Tol)3); which react with tBu2P–P=PMetBu2 to give the phosphinophosphinidene complexes [{η2tBu2P–P}Pt(PMe3)2], [{η2tBu2P–P}Pt(PEt3)2], [{η2tBu2P–P}Pt(PPhEt2)2], [{η2tBu2P–P}Pt(PPh2Et)2], [{η2tBu2P–P}Pt(PPh2Me)2], [{η2tBu2P–P}Pt(PPh2iPr], [{η2tBu2P–P}Pt(PPh2tBu)2] and [{η2tBu2P–P}Pt(P(p‐Tol)3)2]. [{η2tBu2P–P}Pt(PPh3)2] reacts with PMe3 and PEt3 as well as with tBu2PMe, PiPr3 and P(c‐Hex)3 by substituting one PPh3 ligand to give [{η2tBu2P1–P2}Pt(P3Me3)(P4Ph3)], [{η2tBu2P1–P2}Pt(P3Ph3)(P4Me3)], [{η2tBu2P1–P2}Pt(P3Et3)(P4Ph3)], [{η2tBu2P1–P2}Pt(P3MetBu2)(P4Ph3)], [{η2tBu2P1–P2}Pt(P3iPr3)(P4Ph3)] and [{η2tBu2P1–P2}Pt(P3(c‐Hex)3)(P4Ph3)]. With tBu2PMe, [{η2tBu2P–P}Pt(P(p‐Tol)3)2] forms [{η2tBu2P1–P2}Pt(P3MetBu2)(P4(p‐Tol)3)]. The NMR data of the compounds are given and discussed with respect to the influence of the PR3 ligands.  相似文献   

10.
Ternary Thallium Indium Sulfides: A Summary Combined thermal and X-Ray analyses in the ternary system Thallium—Indium—Sulfur show, that the two binary sections Tl2S? In2S3 and TlS? InS contain ternary compounds with unique crystal structures. The chemical formulas of these ternary solids are TlIn5S8, TlIn3S5, TlInS2 and Tl3InS3 for the section Tl2S? In2S3 and TlIn5S6 as well as Tl3In5S8 (metastable high temperature phase) for the section TlS? InS respectively. With TlIn5S7 an additional ternary solid could be detected, which is located outside the two sections. It is derived from the binary mixed valence compound In6S7 by complete substitution of In+ by Tl+. The following ionic formulations make the mixed valence character of the ternary Thallium—Indium-Sulfides reasonable: TlIn5S8 = Tl+(In3+)5(S2?)8, TlIn3S5 = Tl+ (In3+)3(S2?)5, TlInS2 = Tl+In3+(S2?)2, Tl3InS3 = (Tl+)3In3+ · (S2?)3, TlIn5S6 = Tl+([In2]4+)2In3+ (S2?)6, Tl3In5S8 = 4 × [(Tl+)0,75 · (In+)0,25In3+(S2?)2], TlIn5S7 = Tl+[In2]4+ (In3+)3(S2?)7. All compounds contain Tl+-ions in a characteristic “lone pair coordination” of S2? ions. Indium atoms however occur with the oxidation numbers +2 (formal, In2 dumb bells with covalent In? In bonding) and +3 (with In3+ in tetrahedral and octahedral coordination of S2?). Chemical preparation, crystal chemistry and general properties of the ternary solids are discussed, summarized and compared to each other.  相似文献   

11.
The metallation of the η5-C5H5(CO)2Fe-η15-C5H4Mn(CO)3 complex with BunLi (THF, ?78 °C) followed by the treatment of the lithium derivative with Ph2PCl afforded the η5-Ph2PC5H4(CO)2Fe-η15-C5H4Mn(CO)3 complex. The reaction of the latter with η5-C5H5(CO)3WCl in the presence of Me3NO produced the trinuclear complex η5-C5H5Cl(CO)2W-η15-(Ph2P)C5H4(CO)2Fe-η15-C5H4Mn(CO)3. The structure of the latter complex was established by IR, UV, and 1H and 31P NMR spectroscopy and X-ray diffraction. The reaction of MeSiCl3 with three equivalents of LiC5H4(CO)2Fe-η15-C5H4Mn(CO)2PPh3 gave the hexanuclear complex MeSi[C5H4(CO)2Fe-η15-C5H4Mn(CO)2PPh3]3.  相似文献   

12.
Upconversion luminescence tuning of β‐NaYF4 nanorods under 980 nm excitation has successfully been achieved by tridoping with Ln3+ ions with different electronic structures. The effects of Ce3+ ions on NaYF4:Yb3+/Ho3+ as well as Gd3+ ions on NaYF4:Yb3+/Tm3+(Er3+) have been studied in detail. By tridoping with Ce3+ ions, not only were unusual 5G55I7 and 5F2/3K85I8 transitions from Ho3+ ions and 5d→4f transitions from Ce3+ ions observed in NaYF4:Yb3+/Ho3+ nanorods, but also an increase in the intensity of 5F55I8 relative to 5S2/5F45I8 with increasing Ce3+ concentration, which can be attributed to efficient energy transfers of 5I6 (Ho)+2F5/2 (Ce)→5I7 (Ho)+2F7/2 (Ce) and 5S2/5F4 (Ho)+2F5/2 (Ce)→5F5 (Ho)+2F7/2 (Ce). Interestingly, with increasing pump power density, the luminescence of NaYF4:Yb3+/Ho3+ nanorods is always dominated by the 5S2/5F45I8 transition, whereas the luminescence of Ce3+‐tridoped NaYF4:Yb3+/Ho3+ nanorods is dominated by the 5S2/5F45I8 and 5G55I7 transitions in turn. These observations are discussed on the basis of a rate equation model. Furthermore, Gd3+‐tridoped NaYF4:Yb3+/Tm3+(Er3+) nanorods can emit multicolor upconversion emissions spanning from the UV to the near‐infrared under 980 nm excitation. 6P5/28S7/2 (≈306 nm) and 6P7/28S7/2 (≈311 nm) transitions from Gd3+ ions were observed. In addition to the aforementioned luminescence properties, these Gd3+‐tridoped nanorods also exhibit paramagnetic behavior at room temperature and superparamagnetic behavior at 2 or 5 K.  相似文献   

13.
[V2O]+ remains “invisible” in the thermal gas‐phase reaction of bare [V2]+ with CO2 giving rise to [V2O2]+; this is because the [V2O]+ intermediate is being consumed more than 230 times faster than it is generated. However, the fleeting existence of [V2O]+ and its involvement in the [V2]+ → [V2O2]+ chemistry are demonstrated by a cross‐over labeling experiment with a 1:1 mixture of C16O2/C18O2, generating the product ions [V216O2]+, [V216O18O]+, and [V218O2]+ in a 1:2:1 ratio. Density functional theory (DFT) calculations help to understand the remarkable and unexpected reactivity differences of [V2]+ versus [V2O]+ towards CO2.  相似文献   

14.
A diamagnetic AuI4CoIII2 hexanuclear complex, [Au4Co2(dppe)2(l ‐nmc)4]2+ ([ 1L ‐ nmc ]2+; dppe=1,2‐bis(diphenylphosphino)ethane, l ‐H2nmc=N‐methyl‐l ‐cysteine), was newly synthesized by the reaction of [Co(l ‐nmc)2]? with [Au2Cl2(dppe)] and crystallized with different inorganic anions (X=ClO4?, NO3?, Cl?, SO42?) to produce ionic solids ([ 1L ‐ nmc ]Xn). Single‐crystal X‐ray analysis revealed that all the solids crystallize in the chiral space group F432 with a face‐centered‐cubic lattice structure consisting of supramolecular octahedra of complex cations. The paramagnetic nature of all the solids was evidenced by magnetic susceptibility measurements, showing the variation of the oxidation states of two cobalt centers in [ 1L ‐ nmc ]n+ from CoII1.00CoIII1.00 for X=ClO4? or NO3? to CoII0.67CoIII1.33 for X=Cl?, via CoII0.83CoIII1.17 for X=SO42?. The difference in the CoII/III mixed‐valences was explained by the difference in sizes and charges of counter anions accommodated in lattice interstices with a fixed volume.  相似文献   

15.
Preparation of CF3SClF+MF6? (M = As, Sb) and Crystal Structure of CF3SCl2+SbF6? CF3SClF+MF6? (M = As, Sb) is prepared by oxidative fluorination of CF3SCl with XeF+MF6?. The new salt is characterized by IR, Raman and NMR spectra in comparison with CF3SF2+MF6? and CF3SCl2+MF6?. In SO2 solution CF3SClF+SbF6? symmetrizises into CF3SF2+SbF6? and crystalline CF3SCl2+SbF6? with the monoclinic space group P21/c with a = 773.5(14) pm, b = 954.8(15) pm, c = 1242.0(18) pm, β = 100.24(8)°, Z = 4.  相似文献   

16.
采用二阶微扰理论MP2、密度泛函B3LYP方法和含时密度泛函TD-B3LYP方法分别优化了TiO2分子的基态1A1和六个激发态1B23B21B13B11A23A2的几何结构. 1A11B23B21B13B1具有弯曲几何结构, 1A23A2具有线性对称结构. 我们发现激发态1B23B21B13B1键偶极矩的数值大小顺序和相应的键角大小顺序完全一致. 另外, 采用完全活化空间自洽场(CASSCF)CASSCF(6,6)、CASSCF(8,8)、多参考组态相互作用(MRCI)和含时密度泛函TD-B3LYP 计算了TiO2 分子各激发态的垂直激发能和绝热激发能. 对1B23B21B1三个态, MRCI/CASSCF(6,6) 计算的垂直激发能和绝热激发能与已有的实验值最接近. 对其他三个激发态3B11A23A2, 计算的激发能和文献报道的激发能计算值基本一致. 最后, 还计算了TiO2分子的基态和激发态的偶极矩. 对1A11B2态, 偶极矩的计算值与已有的实验值相吻合. 采用原子偶极矩校正的Hirshfeld 布居方法计算了TiO2分子在1A11B23B21B13B1态时各原子的电荷, 发现从基态到激发态偶极矩的变化与电荷从氧原子向钛原子的转移有关. 整个计算中还考察了基函数cc-pVDZ、cc-pVTZ和cc-pVQZ对计算结果的影响.  相似文献   

17.
The 300 K reactions of O2 with C2(X 1Σ+g), C2(a 3 Πu), C3(X? 1Σ+g) and CN(X 2Σ+), which are generated via IR multiple photon dissociation (MPD), are reported. From the spectrally resolved chemiluminescence produced via the IR MPD of C2H3CN in the presence of O2, CO molecules in the a 3Σ+, d 3Δi, and e 3Σ? states were identified, as well as CH(A 2Δ) and CN(B 2Σ+) radicals. Observation of time resolved chemiluminescence reveals that the electronically excited CO molecules are formed via the single-step reactions C2(X 1Σ+g, a 3Πu) + O2 → CO(X 1Σ+ + CO(T), where T denotes are electronically excited triplet state of CO. The rate coefficients for the removal of C2(X 1Σ+g) and C2(a 3Πu) by O2 were determined both from laser induced fluorescence of C2(X 1Σ+g) and C2(a 3Πu), and from the time resolved chemiluminescence from excited CO molecules, and are both (3.0 ± 0.2)10?12 cm3 molec?1 s?1. The rate coefficient of the reaction of C3 with O2, which was determined using the IR MPD of allene as the source of C3 molecules, is <2 × 10?14 cm3 molec?1 s?1. In addition, we find that rate coefficients for C3 reactions with N2, NO, CH4, and C3H6 are all < × 10?14 cm3 molec?1 s?1. Excited CH molecules are produced in a reaction which proceeds with a rate coefficient of (2.6 ± 0.2)10?11 cm3 molec?1 s?1. Possible reactions which may be the source of these radicals are discussed. The reaction of CN with O2 produces NCO in vibrationally excited states. Radiative lifetime of the ā 2Σ state of NCo and the ā 1Πu(000) state of C3 are reported.  相似文献   

18.
Coordination Chemistry of P‐rich Phosphanes and Silylphosphanes. XVII [1] [Co(g5‐Me5C5)(g3tBu2PPCH–CH3)] from [Co(g5‐Me5C5)(g2‐C2H4)2] and tBu2P–P=P(Me)tBu2 [Co(η5‐Me5C5)(η3tBu2PPCH–CH3)] 1 is formed in the reaction of [Co(η5‐Me5C5)(η2‐C2H4)2] 2 with tBu2P–P 4 (generated from tBu2P–P=P(Me)tBu2 3 ) by elimination of one C2H4 ligand and coupling of the phosphinophosphinidene with the second one. The structure of 1 is proven by 31P, 13C, 1H NMR spectra and the X‐ray structure analysis. Within the ligand tBu2P1P2C1H–CH3 in 1 , the angle P1–P2–C1 amounts to 90°. The Co, P1, P2, C1 atoms in 1 look like a „butterfly”︁. The reaction of 2 with a mixture of tBu2P–P=P(Me)tBu2 3 and tBu–C?P 5 yields [Co(η5‐Me5C5){η4‐(tBuCP)2}] 6 and 1 . While 6 is spontaneously formed, 1 appears only after complete consumption of 5 .  相似文献   

19.
The reaction of OH? with O3 eventually leads to the formation of .OH radicals. In the original mechanistic concept (J. Staehelin, J. Hoigné, Environ. Sci. Technol. 1982 , 16, 676–681), it was suggested that the first step occurred by O transfer: OH?+O3→HO2?+O2 and that .OH was generated in the subsequent reaction(s) of HO2? with O3 (the peroxone process). This mechanistic concept has now been revised on the basis of thermokinetic and quantum chemical calculations. A one‐step O transfer such as that mentioned above would require the release of O2 in its excited singlet state (1O2, O2(1Δg)); this state lies 95.5 kJ mol?1 above the triplet ground state (3O2, O2(3Σg?)). The low experimental rate constant of 70 M ?1 s?1 is not incompatible with such a reaction. However, according to our calculations, the reaction of OH? with O3 to form an adduct (OH?+O3→HO4?; ΔG=3.5 kJ mol?1) is a much better candidate for the rate‐determining step as compared with the significantly more endergonic O transfer (ΔG=26.7 kJ mol?1). Hence, we favor this reaction; all the more so as numerous precedents of similar ozone adduct formation are known in the literature. Three potential decay routes of the adduct HO4? have been probed: HO4?→HO2?+1O2 is spin allowed, but markedly endergonic (ΔG=23.2 kJ mol?1). HO4?→HO2?+3O2 is spin forbidden (ΔG=?73.3 kJ mol?1). The decay into radicals, HO4?→HO2.+O2.?, is spin allowed and less endergonic (ΔG=14.8 kJ mol?1) than HO4?→HO2?+1O2. It is thus HO4?→HO2.+O2.? by which HO4? decays. It is noted that a large contribution of the reverse of this reaction, HO2.+O2.?→HO4?, followed by HO4?→HO2?+3O2, now explains why the measured rate of the bimolecular decay of HO2. and O2.? into HO2?+O2 (k=1×108 M ?1 s?1) is below diffusion controlled. Because k for the process HO4?→HO2.+O2.? is much larger than k for the reverse of OH?+O3→HO4?, the forward reaction OH?+O3→HO4? is practically irreversible.  相似文献   

20.
A novel AuICoIII coordination system that is derived from the newly prepared [Co(D ‐nmp)2] ( 1 ; D ‐nmp=N‐methyl‐D ‐penicillaminate) and a gold(I) precursor AuI is reported. Complex 1 acts as a sulfur‐donating metallaligand and reacts with the gold(I) precursor to give [Au2Co2(D ‐nmp)4] ( 2 ), which has an eight‐membered AuI2CoIII2 metallaring. Treatment of 2 with [Au2(dppe)2]2+ (dppe=1,2‐bis(diphenylphosphino)ethane) leads to the formation of [Au4Co2(dppe)2(D ‐nmp)4]2+ ( 3 2+), which consists of an 18‐membered AuI4CoIII2 metallaring that accommodates a tetrahedral anion (BF4, ClO4, ReO4). In solution, the metallaring structure of 3 2+ is readily interconvertible with the nine‐membered AuI2CoIII metallaring structure of [Au2Co(dppe)(D ‐nmp)2]+ ( 4 +); this process depends on external factors, such as solvent, concentration, and nature of the counteranion. These results reveal the lability of the Au S and Au P bonds, which is essential for metallaring expansion and contraction.  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