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.
Zirconium diboride (ZrB2) nanoparticles were synthesized by sol-gel method using zirconium n-propoxide (Zr(OPr)4), boric acid (H3BO3), and sucrose (C12H22O11). Additionally, acetylacetone (acac) was used as chemical modifier in a neutral condition to stabilize Zr(OPr)4 which hydrolyzes easily. Here, C12H22O11 was used since it can be completely decomposed to carbon. Thus, carbon might be accounted precisely for the carbothermal reduction reaction. A single phase ZrB2 without residual ZrO2 was obtained with a molar ratio of B/Zr = 2.3 for the starting materials at 1,550 °C and the average grain size of ZrB2 nanoparticles was ca. 50 nm. The photomicrograph revealed a spherical round shape morphology of the ZrB2 nanoparticles with an uniform size distribution. On the other hand, in the case of either B/Zr (mol.) = 2.0 or pyrolyzing temperature below 1,550 °C for B/Zr (mol.) = 2.3, there existed both m-ZrO2 and t-ZrO2 phases besides ZrB2.  相似文献   

2.
Cr-Al binary compounds are regarded as the potential high-temperature structural materials. However, the structure and important properties of Cr-Al compounds are not completely unclear. Here, we report on the influence of Cr concentration on the structural, mechanical, and thermodynamic properties of Cr-Al compounds by using the first-principles calculations. Four novel Cr-Al compounds, Cr3Al8 with monoclinic structure (C2/m), Cr3Al5 with hexagonal structure (P63mc), Cr2Al3 with tetragonal structure (I4/mmm), and Cr3Al with cubic structure (Pm-3 m), are predicted. The calculated elastic modulus of Cr-Al compounds gradually increases with increasing Cr concentration. Compared to other Cr-Al compounds, our predicted Cr3Al with cubic structure exhibits a strong deformation resistance and high hardness due to symmetrical Cr Al bonds. However, the Debye temperature of Cr7Al3 is larger than that of other Cr-Al compounds. The calculated phonon density of state shows that the high-temperature thermodynamic properties of Cr-Al compounds are attributed to the vibration of Al atom and Cr Al bond.  相似文献   

3.
4.
The B2-type cubic Zr-based compounds are attractive advanced high-temperature materials because of the strong and symmetrical bonds. However, the mechanical and thermodynamic properties of the B2-type cubic Zr-based compounds are not well understood. Here, we use the first-principles calculations to investigate the structural, elastic modulus, ductility, and thermodynamic properties of TMZr (TM = Ru, Mo, Rh, Os, and Re) compounds. Two novel TMZr compounds, MoZr and ReZr, are first predicted by using the phonon dispersion and formation enthalpy, respectively. The results show that the B2-type TMZr compounds not only exhibit high elastic modulus but also show better ductility due to the symmetrical TM-Zr metallic bonds. In particular, the calculated elastic modulus of OsZr is larger than that of the other four TMZr compounds, indicating that the OsZr shows the strongest deformation resistance in five TMZr compounds. The calculated Θ D of RuZr is 328 K, which is larger than that of the other four TMZr compounds. The calculated phonon density of state shows that the high-temperature thermodynamic properties of TMZr derive from the vibration of Zr atom. Therefore, our work predicts that the B2-type OsZr is an attractive high-temperature structural material.  相似文献   

5.
The thermolysis of Zr(BH4)4 vapor at 573 and 623 K in a vacuum of 1.33 × 10−1 Pa was studied. Nanosized zirconium diboride was produced as an X-ray amorphous powder and a crystalline film. According to electron microscopy data, the X-ray amorphous zirconium diboride powder obtained at 573 or 623 K consists of spherical particles 30–40 nm in diameter, which is in quite a good agreement with the equivalent particle diameter (∼35 nm) calculated from the specific surface area of ZrB2. After annealing at 1273 K, the X-ray amorphous zirconium diboride powder crystallizes into a hexagonal lattice with the unit cell parameters a = 0.3159 nm and c = 0.3527 nm. The coherent scattering length D hkl is ∼27 nm. The zirconium diboride film produced at 573 or 623 K crystallizes into a hexagonal lattice with the unit cell parameters a = 0.3163−0.3168 nm and c = 0.3524−0.3531 nm. The coherent scattering length D hkl is ∼14 nm. The thickness of the ZrB2 film on quartz, glass ceramics, and stainless steel is 5–7 μm. The microhardness of the film on a stainless steel substrate under a load of 20 g is 17.8 GPa.  相似文献   

6.
The compound tetramethyl μ-monothiopyrophosphate (C4H12O6P2S) crystallizes in the monoclinic space group C 2/c, with (at -130°C) a = 10.322 Å, b = 8.229 Å, c = 12.062 Å, β = 98.44°, and Dcalc = 1.639 g/mL3 and Z = 4. The crystal structure has been determined by single crystal X-ray diffraction to give a final R value of 0.0329 for 614 independent observed reflections [F˚ > 2.5σ(F˚)]. The sulfur atom resides on a crystallographic two-fold axis. The P S P bond angle is 105.4° and the P S bond lengths are 2.093 Å. The bond angles around phosphorus range from 99.1° to 118.2°. The terminal PO bond is 1.465 Å, and the methoxyl P O bond is about 1.556 Å. The H3C O P bond angle is about 119.5°. Many structural features are interpreted in terms of π-bonding to phosphorus. Comparisons with the structures of pyrophosphate and related compounds indicate that the combined effects of increased acuteness of the P S P bond and the increased length of the P—S bonds lead to an increase of about 0.4 Å in the separation of phosphorus atoms in the sulfur-bridging compound. These facts, together with the weakness of the P S bond, must be taken into account in the interpretation of kinetic data for enzymatic reactions of phosphorothiolates as substrates in place of phosphates.  相似文献   

7.
We apply the first-principles approach to study the structural stability, Vickers hardness, and elastic modulus of ReB12 and RuB12. In particular, we further investigate the influence of high pressure on the structural stability and mechanical properties of ReB12 and RuB12. The calculated results show that ReB12 and RuB12 are thermodynamic stability under high pressure. Here, ReB12 is more thermodynamic stability than that of the RuB12. The calculated Vickers hardness of ReB12 and RuB12 is 16.25 and 16.55 GPa, respectively. It is found that the calculated elastic constants and elastic modulus of ReB12 and RuB12 increase with increasing pressure. In particular, the calculated elastic constants and elastic modulus of ReB12 are larger than that of the RuB12. The calculated electronic structure shows that the high hardness and elastic modulus of ReB12 and RuB12 are attributed to the 3D network B-B covalent bonds.  相似文献   

8.
Synthesis and Structure of Two- and Threenuclear Heterometallic Complexes with Nitrido Bridges between Re and Mo The reaction of ReNCl2(PMe2Ph)3 with MoCl4(NCEt)2 yields the heterometallic threenuclear complex [{(Me2PhP)3(EtCN)ClRe≡N–}2MoCl4][MoNCl5]. The anion [MoNCl5]2– presumably results from a transfer of the nitrido ligand from the Re to the Mo atom. The air-sensitive compound is paramagnetic with μeff = 2.87 B. M. at room temperature. A reduction of the magnetic moment to 1.74 B.M at 20 K starts at 140 K. The complex crystallizes in the orthorhombic space group Pca21 with a = 2430(1), b = 1328(1), c = 2436.3(2) pm, Z = 4. With bond angles Re–N–Mo of 164° and 167° the nitrido bridges are almost linear. The distances Re–N of 169 and 170 pm can be interpreted with triple bonds. The Mo–N bond lengths of 210 and 211 pm correspond to single bonds. In the anion [MoNCl5]2– the distance Mo≡N is 167 pm. Hydrolysis of the threenuclear complex results in a cleavage of one of the nitrido bridges to yield (Me2PhP)3(EtCN)ClRe≡N–MoOCl4. The compound is paramagnetic with μeff = 1.71 B.M. at room temperature. It crystallizes in the orthorhombic space group Pbca with a = 1718.5(4), b = 2037(1), c = 2041.1(7) pm, Z = 8. In the dinuclear complex the [MoOCl4] unit is only weakly coordinated to the nitrido ligand with Mo–N = 246.5 pm, while the distance of the Re≡N bond of 168.1 pm is almost unchanged in comparison with a terminal bond. The bond angle Re≡N–Mo is 165.6°.  相似文献   

9.
Although Ti-V based high-temperature alloys are used in aerospace engine, rocket engine and hot sections, the structure and mechanical properties of Ti-V alloys remains controversy. To explore the correlation between structural and mechanical properties, we apply employed the DFT method to study the phases stability, mechanical and thermodynamic properties of Ti-V solid solution. Two Ti-V solid solutions: Ti(V)ss solid solution and V(Ti)ss solid solution are discussed. Two Ti-V solid solutions are thermodynamic stability. In particular, the Ti-V solid solution prefers to form V(Ti)ss solid solution, in while the V(Ti)ss solid solution remains cubic structure. Furthermore, the Ti(V)ss solid solution is a mechanical instability. However, the V(Ti)ss solid solution is a mechanical stability. Here, the bulk modulus, shear modulus and Young's modulus of V(Ti)ss solid solution are 136.9, 23.5 and 66.7 GPa. In particular, the bulk modulus of V(Ti)ss solid solution is higher than the bulk modulus of the pure Ti. In addition, the V(Ti)ss solid solution shows better ductility compared to the pure Ti and V. Naturally, the stability and mechanical properties of V(Ti) solid solution is related to the Ti-V metallic bond because of the localized hybridization between the Ti(3d) and V(3d).  相似文献   

10.
In the search of the protocovalent bonding, previously recognized in the nitrous acid (HONO), a nature of the chemical bonds in the alkaline nitrites MONO (M = Li, Na, K) has been studied by means of the topological analysis of the Electron Localization Function (ELF) and Electron Localizability Indicator (ELI‐D). Calculations carried out with the B3LYP and MP2(full) methods, in conjunction with the aug‐cc‐pVTZ and 6‐311++G(3df,3pd) basis sets, revealed the cis (C2v, more stable) and trans (Cs) isomers as minima on PES. Alkaline nitrites consist of the alkali metal cation Mδ+ interacting, mainly via electrostatic forces, with the nitrite anion [ONO]δ− (δ ≈ 1e). The covalent N O bonds are characterized by disynaptic basins V(N,O) with the basin populations: 1.58÷1.62e for cis‐Mδ+[ONO]δ− but 1.39÷1.49e for single N O bond and 1.81÷1.87e for formally double NO bond in trans Mδ+[O NO]δ−. The protocovalent nitrogen–oxygen bond has not been observed. The N O bonds are slightly polarized towards the nitrogen atom with the polarity index pNO ≤ 0.12. Two different sets of the hybrid (Lewis) structures are compared leading to different pictures of the bonding. According to NBO data there is a delocalization between the single N O and double NO type bonds, meanwhile results of the ELF analysis emphasize an electron delocalization between the single N O and ionic ON+ hybrids. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

11.
The reaction of [Zr(TrenDMBS)(Cl)] [ Zr1 ; TrenDMBS=N(CH2CH2NSiMe2But )3] with NaPH2 gave the terminal parent phosphanide complex [Zr(TrenDMBS)(PH2)] [ Zr2 ; Zr−P=2.690(2) Å]. Treatment of Zr2 with one equivalent of KCH2C6H5 and two equivalents of benzo‐15‐crown‐5 ether (B15C5) afforded an unprecedented example (outside of matrix isolation) of a structurally authenticated transition‐metal terminal parent phosphinidene complex [Zr(TrenDMBS)(PH)][K(B15C5)2] [ Zr3 ; Zr=P=2.472(2) Å]. DFT calculations reveal a polarized‐covalent Zr=P double bond, with a Mayer bond order of 1.48, and together with IR spectroscopic data also suggest an agostic‐type Zr⋅⋅⋅HP interaction [∡ZrPH=66.7°] which is unexpectedly similar to that found in cryogenic, spectroscopically observed phosphinidene species. Surprisingly, computational data suggest that the Zr=P linkage is similarly polarized, and thus as covalent, as essentially isostructural U=P and Th=P analogues.  相似文献   

12.
The theoretical study of the dehydrogenation of 2,5‐dihydro‐[furan ( 1 ), thiophene ( 2 ), and selenophene ( 3 )] was carried out using ab initio molecular orbital (MO) and density functional theory (DFT) methods at the B3LYP/6‐311G**//B3LYP/6‐311G** and MP2/6‐311G**//B3LYP/6‐311G** levels of theory. Among the used methods in this study, the obtained results show that B3LYP/6‐311G** method is in good agreement with the available experimental values. Based on the optimized ground state geometries using B3LYP/6‐311G** method, the natural bond orbital (NBO) analysis of donor‐acceptor (bond‐antibond) interactions revealed that the stabilization energies associated with the electronic delocalization from non‐bonding lone‐pair orbitals [LP(e)X3] to δ*C(1)  H(2) antibonding orbital, decrease from compounds 1 to 3 . The LP(e)X3→δ*C(1)  H(2) resonance energies for compounds 1 – 3 are 23.37, 16.05 and 12.46 kJ/mol, respectively. Also, the LP(e)X3→δ*C(1)  H(2) delocalizations could fairly explain the decrease of occupancies of LP(e)X3 non‐bonding orbitals in ring of compounds 1 – 3 ( 3 > 2 > 1 ). The electronic delocalization from LP(e)X3 non‐bonding orbitals to δ*C(1)  H(2) antibonding orbital increases the ground state structure stability, Therefore, the decrease of LP(e)X3→δ*C(1)  H(2) delocalizations could fairly explain the kinetic of the dehydrogenation reactions of compounds 1 – 3 (k 1 >k 2 >k 3 ). Also, the donor‐acceptor interactions, as obtained from NBO analysis, revealed that the (C(4)C(7)→δ*C(1)  H(2) resonance energies decrease from compounds 1 to 3 . Further, the results showed that the energy gaps between (C(4)C(7) bonding and δ*C(1)  H(2) antibonding orbitals decrease from compounds 1 to 3 . The results suggest also that in compounds 1 – 3 , the hydrogen eliminations are controlled by LP(e)→δ* resonance energies. Analysis of bond order, natural bond orbital charges, bond indexes, synchronicity parameters, and IRC calculations indicate that these reactions are occurring through a concerted and synchronous six‐membered cyclic transition state type of mechanism.  相似文献   

13.
Density functional theory calculations on α/β-d-mannose (α/β-d-Man) and the corresponding glycosides of n-octyl-α/β-d-mannopyranoside (C8O-α/β-d-Man) were carried out for geometrical optimisation and stability predictions at the B3LYP/6-31G level of theory. These compounds are related anomerically, since they differ by only the orientation of the hydroxyl group at the C1 position. The aim of this study is to investigate the effect of the hydroxyl group's orientations (axial vs. equatorial) at the C1 position on the intra-molecular interactions and the conformational stability of these isomers. The structural parameters of X-H???Y intra-molecular hydrogen bonds were analysed, while the nature of these bonds was considered using the atoms-in-molecules (AIM) approach. Natural bond orbital (NBO) analysis was used to determine bond orders and the effective non-bonding interactions. We have also reported thermodynamic properties and the electronic properties, such as the highest occupied molecular orbital, lowest unoccupied molecular orbital, ionisation energy, electron affinity, electronic chemical potential, chemical hardness, softness and electrophilicity index in the gas phase for all compounds. These results showed that while α-anomers possess only one intra-molecular hydrogen bond, β-anomers possess two intra-molecular hydrogen bonds, which further confirms the anomalous stability of the latter in the self-assembly phenomena.  相似文献   

14.
Multiple bonds between boron and transition metals are known in many borylene (:BR) complexes via metal dπ→BR back‐donation, despite the electron deficiency of boron. An electron‐precise metal–boron triple bond was first observed in BiB2O? [Bi≡B?B≡O]? in which both boron atoms can be viewed as sp‐hybridized and the [B?BO]? fragment is isoelectronic to a carbyne (CR). To search for the first electron‐precise transition‐metal‐boron triple‐bond species, we have produced IrB2O? and ReB2O? and investigated them by photoelectron spectroscopy and quantum‐chemical calculations. The results allow to elucidate the structures and bonding in the two clusters. We find IrB2O? has a closed‐shell bent structure (Cs, 1A′) with BO? coordinated to an Ir≡B unit, (?OB)Ir≡B, whereas ReB2O? is linear (C∞v, 3Σ?) with an electron‐precise Re≡B triple bond, [Re≡B?B≡O]?. The results suggest the intriguing possibility of synthesizing compounds with electron‐precise M≡B triple bonds analogous to classical carbyne systems.  相似文献   

15.
Ab initio and density functional theory‐based calculations are performed to study the structure, stability, and nature of bonding of superhalogen‐supported noble gas (Ng) compounds of the type HNgY where (Ng = Ar‐Rn; Y = BeF3). Here, BeF3 acts as the superhalogen. Calculations show that the HNgBeF3 spontaneously dissociates into product following the dissociation channels: HNgBeF3 → HBeF3 + Ng and HNgBeF3 → Ng + HF + BeF2. The transition states are optimized and the energy barriers are computed to show the metastable behavior of HNgBeF3. HNgBeF3 molecules are kinetically stable with respect to the first dissociation process having energy barriers of 1.0, 5.0, 10.6, and 13.9 kcal/mol for Ar, Kr, Xe, and Rn analogues, respectively, at CCSD(T)/Aug‐cc‐pVTZ level. These calculations suggest that the HXeBeF3 and HRnBeF3 can be shown to be stable up to ∼100 K temperature with a half‐life of ∼102 seconds. The nature of H Ng and two different types of Ng F bonds in HNgBeF3 molecules is explored through the natural bond orbital and electron density analyses. The large Wiberg bond index (WBI) values for the H Ng bond indicate the formation of almost a single bond in between H‐atoms and Ng‐atoms, whereas small WBI values for the two Ng F bonds indicate a noncovalent interaction in between them. The electron density analysis further supports the covalency of the H Ng bond and noncovalent interaction in the two Ng F bonds in HNgBeF3.  相似文献   

16.
Zirconium diboride (ZrB2) nanoparticles were synthesized by sol-gel method using zirconium n-propoxide (Zr(OPr)4), boric acid (H3BO3), sucrose (C12H22O11), and acetic acid (AcOH). Clearly, it was a non-aqueous solution system at the very beginning of the reactions. Here, AcOH was used as both chemical modifier and solvent to control Zr(OPr)4 hydrolysis. Actually, AcOH could dominate the hydrolysis by self-produced water of the chemical propulsion, rather than the help of outer water. C12H22O11 was selected, since it can be completely decomposed to carbon. Thus, carbon might be accounted precisely for the carbothermal reduction reaction. Furthermore, we investigated the influence of the gelation temperature on the morphology of ZrB2 particles. Increasing the gelation temperature, the particle shapes changed from sphere-like particles at 65 °C to a particle chain at 75 °C, and then form rod-like particles at 85 °C. An in-depth HRTEM observation revealed that the nanoparticles of ZrB2 were gradually fused together to evolve into a particle chain, finally into a rod-like shape. These crystalline nature of ZrB2 related to the gelation temperature obeyed the “oriented attachment mechanism” of crystallography.  相似文献   

17.
Herein, we report the employment of the Mo Mo quintuple bonded amidinate complex to stabilize Group 10 metal fragments {(Et3P)2M} (M=Pd, Pt) and give rise to the isolation of the unprecedented δ complexes. X‐ray analysis unambiguously revealed short contacts between Pd or Pt and two Mo atoms and a slight elongation of the Mo Mo quintuple bond in these two compounds. Computational studies show donation of the Mo Mo quintuple‐bond δ electrons to an empty σ orbital on Pd or Pt, and back‐donation from a filled Pd or Pt dπ orbital into the Mo Mo δ* level (LUMO), consistent with the Dewar–Chatt–Duncanson model.  相似文献   

18.
In this work, we have calculated the thermodynamic parameters of the first steps of the free radical and anionic polymerizations of various unsaturated monomers, using ab initio methods of quantum chemistry. The enthalpies and entropies of polymerization were estimated assuming that they correspond to those of the model reaction A  B(p) + HABAH(p′) → HABABAH(p′) where p and p′ stand for the physical state of the considered species. The enthalpies of polymerization were rationalized using the equation ΔH = −ΔΣ NABEAB + SE(A  B) + SE(HABAH) − SE(HABABAH) where NAB is the number of A  B bonds, EAB the corresponding bond energy, − ΔΣ NABEAB the variation of the sum of the bond energy terms, and SE(X) the thermodynamic stabilization energy of compound X. The preferential mode of polymerization of each monomer was derived from the enthalpies of the initiation and initial propagation steps of the two types of polymerization. Thus, we were able to make some comments concerning the feasibility of the polymerization of the monomers under consideration.  相似文献   

19.
The molecular geometries, vibrational properties, and thermodynamic properties of the clusters (Cl2GaN3)n(n=1–4) have been predicted at the B3LYP/6‐311+G* level. The optimized clusters (Cl2GaN3)n (n=2–4) all possess cyclic structures containing Ga Nα Ga linkages. The relationships between geometrical parameters and oligomerization degree n are discussed. The gas‐phase structures of the trimers prefer to exist in boat‐twisting conformation. As for the tetramer, the S4 symmetry structure is the most stable. The infrared spectra are obtained and assigned by vibrational analysis. Thermodynamic properties derived from the infrared spectra on the basis of statistical thermodynamic principles are linearly correlated with the oligomerization degree n as well as the temperature. Meanwhile, thermodynamic analysis of the gas‐phase reaction suggests that the oligomerization is exothermic and favorable under high temperature.  相似文献   

20.
The potential‐energy surfaces of the reactions of dirhodium tetracarboxylate (Rh2II,II) catalyzed nitrene (NR) insertion into C H bonds were examined by a DFT computational study. A pure Becke exchange functional (B88) rather than a hybrid exchange functional (B3, BHandH) was found to be appropriate for the calculation of the energy difference between the singlet and triplet Rh2II,II–NH nitrene species. Rh2II,II–NR1 (R1=(S)‐2‐methyl‐1‐butylformyl) is thermodynamically more favorable with a free energy lower than that of Rh2II,II–N(PhI)R1. The singlet and triplet states of Rh2II,II–NR1 have similar stability. Singlet Rh2II,II–NR1 undergoes a concerted NR insertion into the C H bond with simultaneous formation of the N H and N C bonds during C H bond cleavage; triplet Rh2II,II–NR1 undergoes H atom abstraction to produce a diradical, followed by subsequent bond formation by diradical recombination. The singlet pathway is favored over the triplet in the context of the free energy of activation and leads to the retention of the chirality of the C atom in the NR insertion product. The reactivities of the C H bonds toward the nitrene‐insertion reaction follow the order tertiary>secondary>primary. Relative reaction rates were calculated for the six reaction pathways examined in this work.  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