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.
The development of \begin{document}$\rm{Bi}_2$\end{document}W\begin{document}$\rm{O}_6$\end{document}-based materials has become one of research hotspots due to the increasing demands on high-efficient photocatalyst responding to visible light. In this work, the effect of high energy radiation (\begin{document}$\gamma$\end{document}-ray) on the structure and the photocatalytic activity of \begin{document}$\rm{Bi}_2$\end{document}W\begin{document}$\rm{O}_6$\end{document} nanocrystals was first studied. No morphological change of \begin{document}$\rm{Bi}_2$\end{document}W\begin{document}$\rm{O}_6$\end{document} nanocrystals was observed by SEM under \begin{document}$\gamma$\end{document}-ray radiation. However, the XRD spectra of the irradiated \begin{document}$\rm{Bi}_2$\end{document}W\begin{document}$\rm{O}_6$\end{document} nanocrystals showed the characteristic 2\begin{document}$\theta$\end{document} of (113) plane shifts slightly from 28.37\begin{document}$^{\rm{o}}$\end{document} to 28.45\begin{document}$^{\rm{o}}$\end{document} with the increase of the absorbed dose, confirming the change in the crystal structure of \begin{document}$\rm{Bi}_2$\end{document}W\begin{document}$\rm{O}_6$\end{document}. The XPS results proved the crystal structure change was originated from the generation of oxygen vacancy defects under high-dose radiation. The photocatalytic activity of \begin{document}$\rm{Bi}_2$\end{document}W\begin{document}$\rm{O}_6$\end{document} on the decomposition of methylene blue (MB) in water under visible light increases gradually with the increase of absorbed dose. Moreover, the improved photocatalytic performance of the irradiated \begin{document}$\rm{Bi}_2$\end{document}W\begin{document}$\rm{O}_6$\end{document} nanocrystals remained after three cycles of photocatalysis, indicating a good stability of the created oxygen vacancy defects. This work gives a new simple way to improve photocatalytic performance of \begin{document}$\rm{Bi}_2$\end{document}W\begin{document}$\rm{O}_6$\end{document} through creating oxygen vacancy defects in the crystal structure by \begin{document}$\gamma$\end{document}-ray radiation.  相似文献   

2.
The geometric structures and vibration frequencies of \begin{document}$ para $\end{document}-chlorofluorobenzene (\begin{document}$ p $\end{document}-ClFPh) in the first excited state of neutral and ground state of cation were investigated by resonance-enhanced multiphoton ionization and slow electron velocity-map imaging. The infrared spectrum of S\begin{document}$ _0 $\end{document} state and absorption spectrum for S\begin{document}$ _1 $\end{document}\begin{document}$ \leftarrow $\end{document}S\begin{document}$ _0 $\end{document} transition in \begin{document}$ p $\end{document}-ClFPh were also recorded. Based on the one-color resonant two-photon ionization spectrum and two-color resonant two-photon ionization spectrum, we obtained the adiabatic excited-state energy of \begin{document}$ p $\end{document}-ClFPh as 36302\begin{document}$ \pm $\end{document}4 cm\begin{document}$ ^{-1} $\end{document}. In the two-color resonant two-photon ionization slow electron velocity-map imagin spectra, the accurate adiabatic ionization potential of \begin{document}$ p $\end{document}-ClFPh was extrapolated as 72937\begin{document}$ \pm $\end{document}8 cm\begin{document}$ ^{-1} $\end{document} via threshold ionization measurement. In addition, Franck-Condon simulation was performed to help us confidently ascertain the main vibrational modes in the S\begin{document}$ _1 $\end{document} and D\begin{document}$ _0 $\end{document} states. Furthermore, the mixing of vibrational modes between S\begin{document}$ _0 $\end{document}\begin{document}$ \rightarrow $\end{document}S\begin{document}$ _1 $\end{document} and S\begin{document}$ _1 $\end{document}\begin{document}$ \rightarrow $\end{document}D\begin{document}$ _0 $\end{document} has been analyzed.  相似文献   

3.
The product branching ratio between different products in multichannel reactions is as important as the overall rate of reaction, both in terms of practical applications (\emph{e.g}. models of combustion or atmosphere chemistry) in understanding the fundamental mechanisms of such chemical reactions. A global ground state potential energy surface for the dissociation reaction of deuterated alkyl halide CD\begin{document}$ _3 $\end{document}CH\begin{document}$ _2 $\end{document}F was computed at the CCSD(T)/CBS//B3LYP/aug-cc-pVDZ level of theory for all species. The decomposition of CD\begin{document}$ _3 $\end{document}CH\begin{document}$ _2 $\end{document}F is controversial concerning C\begin{document}$ - $\end{document}F bond dissociation reaction and molecular (HF, DF, H\begin{document}$ _2 $\end{document}, D\begin{document}$ _2 $\end{document}, HD) elimination reaction. Rice-Ramsperger-Kassel-Marcus (RRKM) calculations were applied to compute the rate constants for individual reaction steps and the relative product branching ratios for the dissociation products were calculated using the steady-state approach. At the different energies studied, the RRKM method predicts that the main channel for DF or HF elimination from 1, 2-elimination of CD\begin{document}$ _3 $\end{document}CH\begin{document}$ _2 $\end{document}F is through a four-center transition state, whereas D\begin{document}$ _2 $\end{document} or H\begin{document}$ _2 $\end{document} elimination from 1, 1-elimination of CD\begin{document}$ _3 $\end{document}CH\begin{document}$ _2 $\end{document}F occurs through a direct three-center elimination. At 266, 248, and 193 nm photodissociation, the main product CD\begin{document}$ _2 $\end{document}CH\begin{document}$ _2 $\end{document}+DF branching ratios are computed to be 96.57%, 91.47%, and 48.52%, respectively; however, at 157 nm photodissociation, the product branching ratio is computed to be 16.11%. Based on these transition state structures and energies, the following photodissociation mechanisms are suggested: at 266, 248, 193 nm, CD\begin{document}$ _3 $\end{document}CH\begin{document}$ _2 $\end{document}F\begin{document}$ \rightarrow $\end{document}absorption of a photon\begin{document}$ \rightarrow $\end{document}TS5\begin{document}$ \rightarrow $\end{document}the formation of the major product CD\begin{document}$ _2 $\end{document}CH\begin{document}$ _2 $\end{document}+DF; at 157 nm, CD\begin{document}$ _3 $\end{document}CH\begin{document}$ _2 $\end{document}F\begin{document}$ \rightarrow $\end{document}absorption of a photon\begin{document}$ \rightarrow $\end{document}D/F interchange of TS1\begin{document}$ \rightarrow $\end{document}CDH\begin{document}$ _2 $\end{document}CDF\begin{document}$ \rightarrow $\end{document}H/F interchange of TS2\begin{document}$ \rightarrow $\end{document}CHD\begin{document}$ _2 $\end{document}CHDF\begin{document}$ \rightarrow $\end{document}the formation of the major product CHD\begin{document}$ _2 $\end{document}+CHDF.  相似文献   

4.
The photochemical reaction of potassium ferrocyanide (K\begin{document}$ _4 $\end{document}Fe(CN)\begin{document}$ _6 $\end{document}) exhibits excitation wavelength dependence and non-Kasha rule behavior. In this study, the excited-state dynamics of K\begin{document}$ _4 $\end{document}Fe(CN)\begin{document}$ _6 $\end{document} were studied by transient absorption spectroscopy. Excited state electron detachment (ESED) and photoaquation reactions were clarified by comparing the results of 260, 320, 340, and 350 nm excitations. ESED is the path to generate a hydrated electron (e\begin{document}$ _{\rm{aq}}^{-} $\end{document}). ESED energy barrier varies with the excited state, and it occurs even at the first singlet excited state (\begin{document}$ ^{1} $\end{document}T\begin{document}$ _{\rm{1g}} $\end{document}). The \begin{document}$ ^{1} $\end{document}T\begin{document}$ _{\rm{1g}} $\end{document} state shows \begin{document}$ {\sim} $\end{document}0.2 ps lifetime and converts into triplet [Fe(CN)\begin{document}$ _{6} $\end{document}]\begin{document}$ ^{4-} $\end{document} by intersystem crossing. Subsequently, \begin{document}$ ^{3} $\end{document}[Fe(CN)\begin{document}$ _{5} $\end{document}]\begin{document}$ ^{3-} $\end{document} appears after one CN\begin{document}$ ^{-} $\end{document} ligand is ejected. In sequence, H\begin{document}$ _{2} $\end{document}O attacks [Fe(CN)\begin{document}$ _{5} $\end{document}]\begin{document}$ ^{3-} $\end{document} to generate [Fe(CN)\begin{document}$ _{5} $\end{document}H\begin{document}$ _{2} $\end{document}O]\begin{document}$ ^{3-} $\end{document} with a time constant of approximately 20 ps. The \begin{document}$ ^{1} $\end{document}T\begin{document}$ _{\rm{1g}} $\end{document} state and e\begin{document}$ _{\rm{aq}}^{-} $\end{document} exhibit strong reducing power. The addition of uridine 5\begin{document}$ ' $\end{document}-monophosphate (UMP) to the K\begin{document}$ _{4} $\end{document}Fe(CN)\begin{document}$ _{6} $\end{document} solution decrease the yield of e\begin{document}$ _{\rm{aq}}^{-} $\end{document} and reduce the lifetimes of the e\begin{document}$ _{\rm{aq}}^{-} $\end{document} and \begin{document}$ ^{1} $\end{document}T\begin{document}$ _{\rm{1g}} $\end{document} state. The obtained reaction rate constant of \begin{document}$ ^{1} $\end{document}T\begin{document}$ _{\rm{1g}} $\end{document} state and UMP is 1.7\begin{document}$ {\times} $\end{document}10\begin{document}$ ^{14} $\end{document} (mol/L)\begin{document}$ ^{-1}\cdot $\end{document}s\begin{document}$ ^{-1} $\end{document}, and the e\begin{document}$ _{\rm{aq}}^{-} $\end{document} attachment to UMP is \begin{document}$ {\sim} $\end{document}8\begin{document}$ {\times} $\end{document}10\begin{document}$ ^{9} $\end{document} (mol/L)\begin{document}$ ^{-1}\cdot $\end{document}s\begin{document}$ ^{-1} $\end{document}. Our results indicate that the reductive damage of K\begin{document}$ _{4} $\end{document}Fe(CN)\begin{document}$ _{6} $\end{document} solution to nucleic acids under ultraviolet irradiation cannot be neglected.  相似文献   

5.
Recent detections of C\begin{document}$ _{60} $\end{document}, C\begin{document}$ _{70} $\end{document}, and C\begin{document}$ _{60} $\end{document}\begin{document}$ ^+ $\end{document} in space induced extensive studies of fullerene derivatives in circumstellar environments. As the promising fullerene sources, protoplanetary nebulae (PPNe) shows a number of unidentified bands in their infrared spectra, among which a small sample exhibits an enigmatic feature at \begin{document}$ \sim $\end{document}21 \begin{document}$ \mathtt{μ} $\end{document}m. Hydrogenation converts fullerenes into fulleranes, which breaks the symmetry of fullerene molecules and produces new infrared bands. In this work, we investigate the possibility of fulleranes (C\begin{document}$ _{60} $\end{document}H\begin{document}$ _m $\end{document}) as the carrier of the 21 \begin{document}$ \mathtt{μ} $\end{document}m feature in terms of theoretical vibrational spectra of fulleranes. The evidences favoring and disfavoring the fullerane hypothesis are presented. We made an initial guess for the hydrogen coverage of C\begin{document}$ _{60} $\end{document}H\begin{document}$ _m $\end{document} that may contribute to the 21 \begin{document}$ \mathtt{μ} $\end{document}m feature.  相似文献   

6.
The anionic carbonyl complexes of groups IV and V metals TM(CO)\begin{document}$ _{6,7} $\end{document} (TM=Ti, Zr, Hf, V, Nb, Ta) are prepared in the gas phase using a laser vaporation-supersonic expansion ion source. The infrared spectra of TM(CO)\begin{document}$ _{6,7} $\end{document}\begin{document}$ ^- $\end{document} anion complexes in the carbonyl stretching frequency region are measured by mass-selected infrared photodissociation spectroscopy. The six-coordinated TM(CO)\begin{document}$ _6 $\end{document}\begin{document}$ ^- $\end{document} anions are determined to be the coordination saturate complexes for both the group IV and group V metals. The TM(CO)\begin{document}$ _6 $\end{document}\begin{document}$ ^- $\end{document} complexes of group IV metals (TM=Ti, Zr, Hf) are 17-electron complexes having a \begin{document}$ ^2 $\end{document}A\begin{document}$ _{\rm{1g}} $\end{document} ground state with \begin{document}$ D_{\rm{3d}} $\end{document} symmetry, while the TM(CO)\begin{document}$ _6 $\end{document}\begin{document}$ ^- $\end{document} complexes of group V metals (TM=V, Nb, Ta) are 18-electron species with a closed-shell singlet ground state possessing \begin{document}$ O_{\rm{h}} $\end{document} symmetry. The energy decomposition analyses indicate that the metal-CO covalent bonding is dominated by TM\begin{document}$ ^- $\end{document}(d)\begin{document}$ \rightarrow $\end{document}(CO)\begin{document}$ _6 $\end{document} \begin{document}$ \pi $\end{document}-backdonation and TM\begin{document}$ ^- $\end{document}(d)\begin{document}$ \leftarrow $\end{document}(CO)\begin{document}$ _6 $\end{document} \begin{document}$ \sigma $\end{document}-donation interactions.  相似文献   

7.
In view of the high activity of Pt single atoms in the low-temperature oxidation of CO, we investigate the adsorption behavior of Pt single atoms on reduced rutile TiO\begin{document}$ _2 $\end{document}(110) surface and their interaction with CO and O\begin{document}$ _2 $\end{document} molecules using scanning tunneling microscopy and density function theory calculations. Pt single atoms were prepared on the TiO\begin{document}$ _2 $\end{document}(110) surface at 80 K, showing their preferred adsorption sites at the oxygen vacancies. We characterized the adsorption configurations of CO and O\begin{document}$ _2 $\end{document} molecules separately to the TiO\begin{document}$ _2 $\end{document}-supported Pt single atom samples at 80 K. It is found that the Pt single atoms tend to capture one CO to form Pt-CO complexes, with the CO molecule bonding to the fivefold coordinated Ti (Ti\begin{document}$ _{5 \rm{c}} $\end{document}) atom at the next nearest neighbor site. After annealing the sample from 80 K to 100 K, CO molecules may diffuse, forming another type of complexes, Pt-(CO)\begin{document}$ _2 $\end{document}. For O\begin{document}$ _2 $\end{document} adsorption, each Pt single atom may also capture one O\begin{document}$ _2 $\end{document} molecule, forming Pt-O\begin{document}$ _2 $\end{document} complexes with O\begin{document}$ _2 $\end{document} molecule bonding to either the nearest or the next nearest neighboring Ti\begin{document}$ _{5 \rm{c}} $\end{document} sites. Our study provides the single-molecule-level knowledge of the interaction of CO and O\begin{document}$ _2 $\end{document} with Pt single atoms, which represent the important initial states of the reaction between CO and O\begin{document}$ _2 $\end{document}.  相似文献   

8.
In order to search for high energy density materials, various 4,8-dihydrodifurazano[3,4-b,e]pyrazine based energetic materials were designed. Density functional theory was employed to investigate the relationships between the structures and properties. The calculated results indicated that the properties of these designed compounds were influenced by the energetic groups and heterocyclic substituents. The -N\begin{document}$ _3 $\end{document} energetic group was found to be the most effective substituent to improve the heats of formation of the designed compounds while the tetrazole ring/-C(NO\begin{document}$ _2 $\end{document})\begin{document}$ _3 $\end{document} group contributed much to the values of detonation properties. The analysis of bond orders and bond dissociation energies showed that the addition of -NHNH\begin{document}$ _2 $\end{document}, -NHNO\begin{document}$ _2 $\end{document}, -CH(NO\begin{document}$ _2 $\end{document})\begin{document}$ _3 $\end{document} and -C(NO\begin{document}$ _2 $\end{document})\begin{document}$ _3 $\end{document} groups would decrease the bond dissociation energies remarkably. Compounds A8, B8, C8, D8, E8, and F8 were finally screened as the potential candidates for high energy density materials since these compounds possess excellent detonation properties and acceptable thermal stabilities. Additionally, the electronic structures of the screened compounds were calculated.  相似文献   

9.
Metallophilic interaction is a unique type of weak intermolecular interaction, where the electronic configuration of two metal atoms is closed shell. Despite its significance in multidisciplinary fields, the nature of metallophilic interaction is still not well understood. In this work, we investigated the electronic structures and bonding characteristic of bimetallic Au\begin{document}$ _{2} $\end{document}@Cu\begin{document}$ _{6} $\end{document} nanocluster through density functional theory method, which was reported in experiments recently [Angew. Chem. Int. Ed. 55 , 3611 (2016)]. In general thinking, interaction between two moieties of (CuSH)\begin{document}$ _{6} $\end{document} ring and (Au\begin{document}$ _{2} $\end{document}PH\begin{document}$ _{3} $\end{document})\begin{document}$ _{2} $\end{document} in the Au\begin{document}$ _{2} $\end{document}@Cu\begin{document}$ _{6} $\end{document} nanocluster can be viewed as a d\begin{document}$ ^{10} $\end{document}-\begin{document}$ \sigma $\end{document} closed-shell interaction. However, chemical bonding analysis shows that there is a ten center-two electron (10c-2e) multicenter bonding between two moieties. Further comparative studies on other bimetallic nanocluster M\begin{document}$ _{2} $\end{document}@Cu\begin{document}$ _{6} $\end{document} (M = Ag, Cu, Zn, Cd, Hg) also revealed that multicenter bonding is the origin of electronic stability of the complexes besides the d\begin{document}$ ^{10} $\end{document}-\begin{document}$ \sigma $\end{document} closed-shell interaction. This will provide valuable insights into the understanding of closed-shell interactions.  相似文献   

10.
The structures and electronic properties of the gaseous M\begin{document}$ _2 $\end{document}Pt\begin{document}$ _2 $\end{document}\begin{document}$ ^{0/-} $\end{document} clusters (M represents the alkaline earth metal) were investigated using the density functional theory (B3LYP and PBE0) and wave function theory (SCS-MP2, CCSD and CCSD (T)). The results indicate that the D\begin{document}$ _{2{h}} $\end{document} isomers with the planar structures are more stable than the C\begin{document}$ _{2v} $\end{document} isomers with smaller dihedral angles and shorter Pt-Pt bond lengths. The mutual competition of M(s, p)-Pt(5d) interaction and Pt-Pt covalent bonding contributes to the different stabilizations of the two kinds of isomers. The M(s, p)-Pt(5d) interaction favors the planar isomers with D\begin{document}$ _{2h} $\end{document} symmetry, while the Pt-Pt covalent bonding leads to the C\begin{document}$ _{2v} $\end{document} isomers with bending structures. Two different crossing points are determined in the potential energy curves of Be\begin{document}$ _2 $\end{document}Pt\begin{document}$ _2 $\end{document} with the singlet and triplet states. But there is just one crossing point in potential energy curves of Ra\begin{document}$ _2 $\end{document}Pt\begin{document}$ _2 $\end{document} and Ca\begin{document}$ _2 $\end{document}Pt\begin{document}$ _2 $\end{document}\begin{document}$ ^- $\end{document} because of flatter potential energy curves of Ra\begin{document}$ _2 $\end{document}Pt\begin{document}$ _2 $\end{document} with the triplet state or Ca\begin{document}$ _2 $\end{document}Pt\begin{document}$ _2 $\end{document}\begin{document}$ ^- $\end{document} with quartet state. The results reveal a unique example of dihedral angle-bending isomers with the smallest number of atoms and may help the understanding of the bonding properties of other potential angle-bending isomers.  相似文献   

11.
Poly(3, 4-ethylenedioxythiophene): poly(styrenesulfonate) (PEDOT: PSS) has been widely adopted as hole transport material (HTM) in inverted perovskite solar cells (PSCs), due to high optical transparency, good mechanical flexibility, and high thermal stability; however, its acidity and hygroscopicity inevitably hamper the long-term stability of the PSCs and its energy level does not match well with perovskite materials with a relatively low open-circuit voltage. In this work, p-type delafossite CuCrO\begin{document}$ _2 $\end{document} nanoparticles synthesized through hydrothermal method was employed as an alternative HTM for triple cation perovskite [(FAPbI\begin{document}$ _3 $\end{document})\begin{document}$ _{0.87} $\end{document}(MAPbBr\begin{document}$ _3 $\end{document})\begin{document}$ _{0.13} $\end{document}]\begin{document}$ _{0.92} $\end{document}(CsPbI\begin{document}$ _3 $\end{document})\begin{document}$ _{0.08} $\end{document} (possessing better photovoltaic performance and stability than conventional CH\begin{document}$ _3 $\end{document}NH\begin{document}$ _3 $\end{document}PbI\begin{document}$ _3 $\end{document}) based inverted PSCs. The average open-circuit voltage of PSCs increases from 908 mV of the devices with PEDOT: PSS HTM to 1020 mV of the devices with CuCrO\begin{document}$ _2 $\end{document} HTM. Ultraviolet photoemission spectroscopy demonstrates the energy band alignment between CuCrO\begin{document}$ _2 $\end{document} and perovskite is better than that between PEDOT: PSS and perovskite, the electrochemical impedance spectroscopy indicates CuCrO\begin{document}$ _2 $\end{document}-based PSCs exhibit larger recombination resistance and longer charge carrier lifetime than PEDOT: PSS-based PSCs, which contributes to the high \begin{document}$ V_{\rm{OC}} $\end{document} of CuCrO\begin{document}$ _2 $\end{document} HTM-based PSCs.  相似文献   

12.
We report a study on photo-ionization of benzene and aniline with incidental subsequent dissociation by the customized reflection time-of-flight mass spectrometer utilizing a deep ultraviolet 177.3 nm laser. Highly efficient ionization of benzene is observed with a weak C\begin{document}$ _4 $\end{document}H\begin{document}$ _3 $\end{document}\begin{document}$ ^+ $\end{document} fragment formed by undergoing disproportional C\begin{document}$ - $\end{document}C bond dissociation. In comparison, a major C\begin{document}$ _5 $\end{document}H\begin{document}$ _6 $\end{document}\begin{document}$ ^{+\cdot} $\end{document} fragment and a minor C\begin{document}$ _6 $\end{document}H\begin{document}$ _6 $\end{document}\begin{document}$ ^{+\cdot} $\end{document} radical are produced in the ionization of aniline pertaining to the removal of CNH\begin{document}$ ^\cdot $\end{document} and NH\begin{document}$ ^\cdot $\end{document} radicals, respectively. First-principles calculation is employed to reveal the photo-dissociation pathways of these two molecules having a structural difference of just an amino group. It is demonstrated that hydrogen atom transfer plays an important role in the cleavage of C\begin{document}$ - $\end{document}C or C\begin{document}$ - $\end{document}N bonds in benzene and aniline ions. This study is helpful to understand the underlying mechanisms of chemical bond fracture of benzene ring and related aromatic molecules.  相似文献   

13.
Photocatalytic degradation of organic pollutants has become a hot research topic because of its low energy consumption and environmental-friendly characteristics. Bismuth oxide (Bi\begin{document}$ _2 $\end{document}O\begin{document}$ _3 $\end{document}) nanocrystals with a bandgap ranging from 2.0 eV to 2.8 eV have attracted increasing attention due to high activity of photodegradation of organic pollutants by utilizing visible light. Though several methods have been developed to prepare Bi\begin{document}$ _2 $\end{document}O\begin{document}$ _3 $\end{document}-based semiconductor materials over recent years, it is still difficult to prepare highly active Bi\begin{document}$ _2 $\end{document}O\begin{document}$ _3 $\end{document} catalysts in large scale with a simple method. Therefore, developing simple and feasible methods for the preparation of Bi\begin{document}$ _2 $\end{document}O\begin{document}$ _3 $\end{document} nanocrystals in large scale is important for the potential applications in industrial wastewater treatment. In this work, we successfully prepared porous Bi\begin{document}$ _2 $\end{document}O\begin{document}$ _3 $\end{document} in large scale via etching commercial BiSn powders, followed by thermal treatment with air. The acquired porous Bi\begin{document}$ _2 $\end{document}O\begin{document}$ _3 $\end{document} exhibited excellent activity and stability in photocatalytic degradation of methylene blue. Further investigation of the mechanism witnessed that the suitable band structure of porous Bi\begin{document}$ _2 $\end{document}O\begin{document}$ _3 $\end{document} allowed the generation of reactive oxygen species, such as O\begin{document}$ _2 $\end{document}\begin{document}$ ^{-\cdot} $\end{document} and \begin{document}$ \cdot $\end{document}OH, which effectively degraded MB.  相似文献   

14.
The kinetics for hydrogen (H) adsorption on Ir(111) electrode has been studied in both HClO\begin{document}$ _4 $\end{document} and H\begin{document}$ _2 $\end{document}SO\begin{document}$ _4 $\end{document} solutions by impedance spectroscopy. In HClO\begin{document}$ _4 $\end{document}, the adsorption rate for H adsorption on Ir(111) increases from 1.74\begin{document}$ \times $\end{document}10\begin{document}$ ^{-8} $\end{document} mol\begin{document}$ \cdot $\end{document}cm\begin{document}$ ^{-2} $\end{document}\begin{document}$ \cdot $\end{document}s\begin{document}$ ^{-1} $\end{document} to 3.47\begin{document}$ \times $\end{document}10\begin{document}$ ^{-7} $\end{document} mol\begin{document}$ \cdot $\end{document}cm\begin{document}$ ^{-2} $\end{document}\begin{document}$ \cdot $\end{document}s\begin{document}$ ^{-1} $\end{document} with the decrease of the applied potential from 0.2 V to 0.1 V (vs. RHE), which is ca. one to two orders of magnitude slower than that on Pt(111) under otherwise identical condition. This is explained by the stronger binding of water to Ir(111), which needs a higher barrier to reorient during the under potential deposition of H from hydronium within the hydrogen bonded water network. In H\begin{document}$ _2 $\end{document}SO\begin{document}$ _4 $\end{document}, the adsorption potential is ca. 200 mV negatively shifted, accompanied by a decrease of adsorption rate by up to one order of magnitude, which is explained by the hindrance of the strongly adsorbed sulfate/bisulfate on Ir(111). Our results demonstrate that under electrochemical environment, H adsorption is strongly affected by the accompanying displacement and reorientation of water molecules that initially stay close to the electrode surface.  相似文献   

15.
A fundamental study on C-C coupling, that is the crucial step in the Fischer-Tropsch synthesis (FTS) process to obtain multi-carbon products, is of great importance to tailor catalysts and then guide a more promising pathway. It has been demonstrated that the coupling of CO with the metal carbide can represent the early stage in the FTS process, while the related mechanism is elusive. Herein, the reactions of the CuC\begin{document}$ _3 $\end{document}H\begin{document}$ ^- $\end{document} and CuC\begin{document}$ _3 $\end{document}\begin{document}$ ^- $\end{document} cluster anions with CO have been studied by using mass spectrometry and theoretical calculations. The experimental results showed that the coupling of CO with the C\begin{document}$ _3 $\end{document}H\begin{document}$ ^- $\end{document} moiety of CuC\begin{document}$ _3 $\end{document}H\begin{document}$ ^- $\end{document} can generate the exclusive ion product COC\begin{document}$ _3 $\end{document}H\begin{document}$ ^- $\end{document}. The reactivity and selectivity of this reaction of CuC\begin{document}$ _3 $\end{document}H\begin{document}$ ^- $\end{document} with CO are greatly higher than that of the reaction of CuC\begin{document}$ _3 $\end{document}\begin{document}$ ^- $\end{document} with CO, and this H-assisted C-C coupling process was rationalized by theoretical calculations.  相似文献   

16.

The simple homodinuclear M

\begin{document}$ - $\end{document}

M single bonds for group II and XII elements are difficult to obtain as a result of the fulfilled s

\begin{document}$ ^2 $\end{document}

electronic configurations, consequently, a dicationic prototype is often utilized to design the M

\begin{document}$ ^+ $\end{document}\begin{document}$ - $\end{document}

M

\begin{document}$ ^+ $\end{document}

single bond. Existing studies generally use sterically bulky organic ligands L

\begin{document}$ ^- $\end{document}

to synthesize the compounds in the L

\begin{document}$ ^- $\end{document}\begin{document}$ - $\end{document}

M

\begin{document}$ ^+ $\end{document}\begin{document}$ - $\end{document}

M

\begin{document}$ ^+ $\end{document}\begin{document}$ - $\end{document}

L

\begin{document}$ ^- $\end{document}

manner. However, here we report the design of Mg

\begin{document}$ - $\end{document}

Mg and Zn

\begin{document}$ - $\end{document}

Zn single bonds in two ligandless clusters, Mg

\begin{document}$ _2 $\end{document}

B

\begin{document}$ _7 $\end{document}\begin{document}$ ^- $\end{document}

and Zn

\begin{document}$ _2 $\end{document}

B

\begin{document}$ _7 $\end{document}\begin{document}$ ^- $\end{document}

, using density functional theory methods. The global minima of both of the clusters are in the form of M

\begin{document}$ _2 $\end{document}\begin{document}$ ^{2+} $\end{document}

(B

\begin{document}$ _7 $\end{document}\begin{document}$ ^{3-} $\end{document}

), where the M

\begin{document}$ - $\end{document}

M single bonds are positioned above a quasi-planar hexagonal B

\begin{document}$ _7 $\end{document}

moiety. Chemical bonding analyses further confirm the existence of Mg

\begin{document}$ - $\end{document}

Mg and Zn

\begin{document}$ - $\end{document}

Zn single bonds in these clusters, which are driven by the unusually stable B

\begin{document}$ _7 $\end{document}\begin{document}$ ^{3-} $\end{document}

moiety that is both

\begin{document}$ \sigma $\end{document}

and

\begin{document}$ \pi $\end{document}

aromatic. Vertical detachment energies of Mg

\begin{document}$ _2 $\end{document}

B

\begin{document}$ _7 $\end{document}\begin{document}$ ^- $\end{document}

and Zn

\begin{document}$ _2 $\end{document}

B

\begin{document}$ _7 $\end{document}\begin{document}$ ^- $\end{document}

are calculated to be 2.79 eV and 2.94 eV, respectively, for the future comparisons with experimental data.

  相似文献   

17.
The structures, energetics, and infrared (IR) spectra of the cationic monomethylamine-water clusters, [(CH\begin{document}$_3$\end{document}NH\begin{document}$_2$\end{document})(H\begin{document}$_2$\end{document}O)\begin{document}$_n$\end{document}]\begin{document}$^+$\end{document} (\begin{document}$n$\end{document}=1\begin{document}$-$\end{document}5), have been studied using quantum chemical calculations at the MP2/6-311+G(2d,p) level. The results reveal that the formation of proton-transferred CH\begin{document}$_2$\end{document}NH\begin{document}$_3$\end{document}\begin{document}$^+$\end{document} ion core structure is preferred via the intramolecular proton transfer from the methyl group to the nitrogen atom and the water molecules act as the acceptor for the O\begin{document}$\cdots$\end{document}HN hydrogen bonds with the positively charged NH\begin{document}$_3$\end{document}\begin{document}$^+$\end{document} moiety of CH\begin{document}$_2$\end{document}NH\begin{document}$_3$\end{document}\begin{document}$^+$\end{document}, whose motif is retained in the larger clusters. The CH\begin{document}$_3$\end{document}NH\begin{document}$_2$\end{document}\begin{document}$^+$\end{document} ion core structure is predicted to be less energetically favorable. Vibrational frequencies of CH stretches, hydrogen-bonded and free NH stretches, and hydrogen-bonded OH stretches in the calculated IR spectra of the CH\begin{document}$_2$\end{document}NH\begin{document}$_3$\end{document}\begin{document}$^+$\end{document} and CH\begin{document}$_3$\end{document}NH\begin{document}$_2$\end{document}\begin{document}$^+$\end{document} type structures are different from each other, which would afford the sensitive probes for fundamental understanding of hydrogen bonding networks generated from the radiation-induced chemical processes in the [(CH\begin{document}$_3$\end{document}NH\begin{document}$_2$\end{document})(H\begin{document}$_2$\end{document}O)\begin{document}$_n$\end{document}]\begin{document}$^+$\end{document} complexes.  相似文献   

18.
The H+CH\begin{document}$ _3 $\end{document}OH reaction, which plays an important role in combustion and the interstellar medium, presents a prototypical system with multiple channels. In this work, mode specific dynamics of different product channels is investigated theoretically on a recently developed reliable potential energy surface based on a large number of data points calculated at the level of UCCSD(T)-F12a/AVTZ. It has been demonstrated that vibrational excitations of the O\begin{document}$ - $\end{document}H stretching motion, the torsional motion, the C\begin{document}$ - $\end{document}H stretching vibrations, show different influences on the four product channels, H\begin{document}$ _2 $\end{document}+CH\begin{document}$ _3 $\end{document}O, H\begin{document}$ _2 $\end{document}+CH\begin{document}$ _2 $\end{document}OH, H\begin{document}$ _2 $\end{document}O+CH\begin{document}$ _3 $\end{document}, and H+CH\begin{document}$ _3 $\end{document}OH. This work is helpful for understanding the mode-specific dynamics and controlling the competition for complicated reactions with multiple product channels.  相似文献   

19.
We constructed two types of copper-doped metal-organic framework (MOF), i.e., Cu@UiO-66-NH\begin{document}$ _2 $\end{document} and Cu-UiO-66-NH\begin{document}$ _2 $\end{document}. In the former, Cu\begin{document}$ ^{2+} $\end{document} ions are impregnated in the pore space of the amine-functionalized, Zr-based UiO-66-NH\begin{document}$ _2 $\end{document}; while in the latter, Cu\begin{document}$ ^{2+} $\end{document} ions are incorporated to form a bimetal-center MOF, with Zr\begin{document}$ ^{4+} $\end{document} being partially replaced by Cu\begin{document}$ ^{2+} $\end{document} in the Zr\begin{document}$ - $\end{document}O oxo-clusters. Ultrafast spectroscopy revealed that the photoinduced relaxation kinetics associated with the ligand-to-cluster charge-transfer state is promoted for both Cu-doped MOFs relative to undoped one, but in a sequence of Cu-UiO-66-NH\begin{document}$ _2 $\end{document}\begin{document}$ > $\end{document}Cu@UiO-66-NH\begin{document}$ _2 $\end{document}\begin{document}$ > $\end{document}UiO-66-NH\begin{document}$ _2 $\end{document}. Such a sequence turned to be in line with the trend observed in the visible-light photocatalytic hydrogen evolution activity tests on the three MOFs. These findings highlighted the subtle effect of copper-doping location in this Zr-based MOF system, further suggesting that rational engineering of the specific metal-doping location in alike MOF systems to promote the photoinduced charge separation and hence suppress the detrimental charge recombination therein is beneficial for achieving improved performances in MOF-based photocatalysis.  相似文献   

20.
The \begin{document}$ C^2\Pi $\end{document}-\begin{document}$ X^2\Pi $\end{document}(0, 0) band of AgO has been reinvestigated by laser induced fluorescence spectroscopy with a spectral resolution of \begin{document}$ \sim $\end{document}0.02 cm\begin{document}$ ^{-1} $\end{document}. The AgO molecules are produced by discharging a gas mixture of O\begin{document}$ _2 $\end{document}/Ar with silver needle electrodes in a supersonic jet expansion. By employing a home-made narrowband single longitude mode optical parametric oscillator (SLM-OPO) as the laser source, high-resolution spectra of the \begin{document}$ C^2\Pi $\end{document}-\begin{document}$ X^2\Pi $\end{document}(0, 0) band have been recorded for both \begin{document}$ ^{107} $\end{document}Ag\begin{document}$ ^{16} $\end{document}O and \begin{document}$ ^{109} $\end{document}Ag\begin{document}$ ^{16} $\end{document}O isotopologues. The spectroscopic constants of the \begin{document}$ C^2\Pi $\end{document} state are consequently determined, with the \begin{document}$ ^{109} $\end{document}Ag\begin{document}$ ^{16} $\end{document}O one being reported for the first time. The nature of the spin-orbit coupling effect in the \begin{document}$ C^2\Pi $\end{document} state is proposed to be due to state mixing with the nearby repulsive \begin{document}$ ^{4}\Sigma^{-} $\end{document} and \begin{document}$ ^{4}\Pi $\end{document} states.  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