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.
To identify thermoelectric materials containing abundant, low-cost and non-toxicelements, we have studied the electronic structures and thermoelectric properties of(Mg2X)2/(Mg2Y)2 (X, Y = Si, Ge, Sn) superlattices withstate-of-the-art first-principles calculations using a modified Becke and Johnson (mBJ)exchange potential. Our results show that (Mg2Ge)2/ (Mg2Sn)2 and(Mg2Si)2/(Mg2Sn)2 are semi-metals using mBJ plusspin-orbit coupling (mBJ +SOC), while (Mg2Si)2/ (Mg2Ge)2 ispredicted to be a direct-gap semiconductor with a mBJ gap value of 0.46 eV andmBJ + SOC gap valueof 0.44 eV. Thermoelectric properties are predicted by through solving the Boltzmanntransport equations within the constant scattering time approximation. It is found that(Mg2Si)2/(Mg2Ge)2 has a larger Seebeck coefficient andpower factor than (Mg2Ge)2/ (Mg2Sn)2 and(Mg2Si)2/(Mg2Sn)2 for both p-type and n-type doping. Thedetrimental influence of SOC on the power factor of p-type (Mg2X)2/(Mg2Y)2 (X, Y = Si, Ge, Sn) is analyzed as afunction of the carrier concentration, but there is a negligible SOC effect for n-type.These results can be explained by the influence of SOC on their valence and conductionbands near the Fermi level.  相似文献   

2.
ABSTRACT

QM(UB3LYP)/MM(AMBER) calculations were performed for the locations of the transition structure (TS) of the oxygen–oxygen (O–O) bond formation in the S4 state of the oxygen-evolving complex (OEC) of photosystem II (PSII). The natural orbital (NO) analysis of the broken-symmetry (BS) solutions was also performed to elucidate the nature of the chemical bonds at TS on the basis of several chemical indices defined by the occupation numbers of NO. The computational results revealed a concerted bond switching (CBS) mechanism for the oxygen–oxygen bond formation coupled with the one-electron transfer (OET) for water oxidation in OEC of PSII. The orbital interaction between the σ-HOMO of the Mn(IV)4–O(5) bond and the π*-LUMO of the Mn(V)1=O(6) bond plays an important role for the concerted O–O bond formation for water oxidation in the CaMn4O6 cluster of OEC of PSII. One electron transfer (OET) from the π-HOMO of the Mn(V)1=O(6) bond to the σ*-LUMO of the Mn(IV)4–O(5) bond occurs for the formation of electron transfer diradical, where the generated anion radical [Mn(IV)4–O(5)]-? part is relaxed to the ?Mn(III)4?…?O(5)- structure and the cation radical [O(6)=Mn(V)1]+ ? part is relaxed to the +O(6)–Mn(IV)1? structure because of the charge-spin separation for the electron-and hole-doped Mn–oxo bonds. Therefore, the local spins are responsible for the one-electron reductions of Mn(IV)4->Mn(III)4 and Mn(V)1->Mn(IV)1. On the other hand, the O(5)- and O(6)+ sites generated undergo the O–O bond formation in the CaMn4O6 cluster. The Ca(II) ion in the cubane- skeleton of the CaMn4O6 cluster assists the above orbital interactions by the lowering of the orbital energy levels of π*-LUMO of Mn(V)1=O(6) and σ*-LUMO of Mn(IV)4–O(5), indicating an important role of its Lewis acidity. Present CBS mechanism for the O–O bond formation coupled with one electron reductions of the high-valent Mn ions is different from the conventional radical coupling (RC) and acid-base (AB) mechanisms for water oxidation in artificial and native photosynthesis systems. The proton-coupled electron transfer (PC-OET) mechanism for the O–O bond formation is also touched in relation to the CBS-OET mechanism.  相似文献   

3.
The reaction of trans-(Co)3Fe(PPh2Py)2 with MeXn gave nine new compounds of trans-(CO)3Fe(PPh2Py)2MeXn (MeXn=Co(SCN)2, Ni(SCN)2, Fe(SCN)3, Cd(SCN)2, Mn(SCN)2, Zn(SCN)2, Mo(CO)3, Hg(SCN)2 and CuBr). The compositions of the compounds were determined through elemental analysis. The structural determination was made by IR, FD-MS and31P-NMR. Mössbauer spectra were taken at 78 K. The solid-state structure of the complex has been determined by a single-crystal X-ray diffraction study. The crystal data for trans-(CO)3Fe(Ph2PPy)2Hg(SCN)2 are:a=16.369(5) Å,b=13.754(3) Å,c=17.749(2) Å,r=108.95(2)°, monoclinic space group P21/n,Z= 4. The determination of ESCA demonstrated the change in the Fe value. In the present paper, the effect of the ligands on Fe-metal bonds is discussed.  相似文献   

4.
Electron spin phase memory times,T m, as a function of temperature were measured for Cu(II) bis(diethyldithiocarbamate), Cu(Et2dtc)2; Cu(II) bis(diethyldithiophosphate), Cu(Et2dtp)2; Cu(II) bis(diphenyldithiophosphate), Cu(Ph2dtp)2; Cu(II) tetratolylporphyrin, CuTTP; vanadyl 5-(4-carboxyphenyl)-10,15,20-tritolylporphyrin, VOTTP-COOH; and Ag(II) tetratolylporphyrin, AgTTP, doped into powdered samples of closely-related diamagnetic hosts. For the three metalloporphyrins, the electron spin relaxation rate (1/T m) increased monotonically with increasing temperature. However the temperature dependence of the relaxation rate was not monotonic for the three other Cu(II) complexes. For Cu(Et2dtc)2 and Cu(Et2dtp)2 the temperature dependence of 1/T m between about 85 and 130 K is attributed to the effects of methyl rotation, with activation energies of 1.0 kcal/mole. Between 120 and 250 K the 1/T m data for Cu(Ph2dtp)2 exhibit effects that are attributed to motion of the phenyl rings.  相似文献   

5.
57Fe Mössbauer spectroscopy, magnetic susceptibility and powder x-ray-diffraction measurements were used to study superconductivity and magnetic order in YBa2(Cu1?xFex)4O8+δ. Tc is decreasing with x, disappearing for x>xc≈0.04. For xc iron substitutes Cu, predominantly in the Cu(1) site exhibiting a single quadrupole Mössbauer spectrum at 90 K. For x>xc magnetic order is observed in the Cu(2) site, TN=380 (5) K for x=0.1 and Heff (Cu(2), 4.2 K)=510(2) kOe. However, the most surprising discovery is that for x=0.025, for which Tc=27(2) K, the Fe in the Cu(1) site orders magnetically at TN=30(2) K and Heff (Cu(1), 4.2 K)=461(2) kOe. The coexistence and competition between superconductivity and magnetic order in the Cu(1) and Cu(2) sites in YBa2Cu4O8 are discussed in terms of the previously observed phase diagrams for Y1?xPrxBa2(Cu1?yFey)3Oz.  相似文献   

6.
Nanostructures of three Zinc(II) coordination polymers, [Zn(NNO)2(H2O)4]n (1), [Zn(PNNO)2(H2O)2]n (2) and [Zn(H2O)6]·(INNO)2 (3) {NNO: Nicotinic acid N-oxide, PNNO: Picolinic acid N-oxide and INNO: Isonicotinic acid N-oxide}, have been synthesized by a sonochemical process and reaction of ligands with Zn(CH3COO)2. The Zinc(II) oxide nano-particles have been synthesized from thermolysis of [Zn(NNO)2(H2O)4]n (1), [Zn(PNNO)2(H2O)2]n (2) and [Zn(H2O)6]·(INNO)2 (3) at two different methods (with surfactant and without surfactant) and two temperatures (200 and 600 °C). The ZnO nanoparticles were characterized by X-ray diffraction (XRD) and scanning electron microscopy (SEM). Comparison of the SEM images of ZnO nano-particles at two different methods and temperatures shows that higher temperature results in an increasing of agglomeration and thus small and spherical ZnO particles with good separation were produced by thermolysis of compounds at 200 °C and by use of surfactant.  相似文献   

7.
Crystal field parameters for Pr3+ in {[Ni(salen)Pr-(hfac)3](H2O)} (noted as NiPr) and {[Ni(salen)Pr(hfac)3(pyr)]-(CHCl3)} (noted as NiPrpyr) have been found from a fit to the thermal variation in the magnetic susceptibility of NiPr and NiPrpyr. The nature of exchange interaction in [Cu(salen)Pr(hfac)3(pyr)] (noted as CuPrpyr), {[{Cu(salen)Pr(hfac)3}2(pyz)](H2O)3} (noted as Cu2Pr2pyz) and {[{Cu(salen)Pr(hfac)3}2(bpy)]-(CHCl3)2} (noted as Cu2Pr2bpy, bpy=4,4_-bipyridine) have been found using the derived results for NiPr and NiPrpyr. All the exchange interactions give significant contribution to the thermal variation in magnetic susceptibility below 50 K. The contribution due to Pr-Cu interaction is positive while that of the Cu-Cu and Pr-Pr interactions are negative. The behaviors below 10 K for Pr-Cu and Pr-Pr are difficult to explain, and point to a possible change in structure of CuPrpyr, Cu2Pr2pyz and Cu2Pr2bpy below 10 K. The theoretical thermal variations in the magnetic specific heat of NiPr and NiPrpyr are computed and discussed.  相似文献   

8.
In this article, the isomerisation mechanisms of HN(NO2)2 to O2NNN(O)OH without and with catalyst X (X = H2O, (H2O)2, (H2O)3, HCOOH, H2SO4, CH3CH2COOH and HN(NO2)2) have been investigated theoretically at the CBS-QB3 level of theory. Our results show that the catalyst X (X = H2O, (H2O)2, (H2O)3, HCOOH, H2SO4 and CH3CH2COOH) shows different positive catalytic effects on reducing the apparent activation energy of the isomerisation reaction processes. Such different catalytic effects are mainly related to the number of hydrogen bonds and the size of the ring structure in X (X = H2O, (H2O)2 and (H2O)3)-assisted transition states, as well as different values of pKa for H2SO4, HCOOH and CH3CH2COOH. Very interesting is also the fact that H2SO4-assisted reaction is the most favourable for the hydrogen transfer from HN(NO2)2 to O2NNN(O)OH, due to the smallest pKa (?3.0) value of H2SO4 than H2O, HCOOH, H2SO4 and CH3CH2COOH, and also because of the largest ∠X???H???Y (the angle between the hydrogen bond donor and acceptor) involved in H2SO4-assisted transition state. Compared to the self-catalysis of the isomerisation mechanisms of HN(NO2)2 to O2NNN(O)OH, the apparent activation energy of H2SO4-assisted channel also reduces by 9.6 kcal?mol?1, indicating that H2SO4 can affect the isomerisation of HN(NO2)2 to O2NNN(O)OH, most obvious among all the catalysts H2O, (H2O)2, (H2O)3, HCOOH, H2SO4, CH3CH2COOH and HN(NO2)2.  相似文献   

9.
The infrared spectra of the complexes [Cu(bipy)3] (C104)2′ [Cu(bipy)2] (C104)2, [Cu(bipy)2I]I, [cu(bipy)2I] C104 and [Cu(bipy)(C104)2] (bipy - 2,2′-bipyridine) are discussed in relation to their known or probable structures. Evidence for tetragonal distortion of [Cu(bipy)3]2+ but not [Cu(phen)3]2+ (phen = 1, 10-phenanthroline) is adduced.  相似文献   

10.
The preparation and photophysics of organometallic Ir(III) complexes with 3-phenylpyridazine (H6Clppdz) and its derivative are reported. IrCl3 · nH2O reacts with H6Clppdz, and 3-chloro-6-(3-methoxyphenyl)-4-methyl-pyridazine (MCOMppdz) to give the corresponding chloride-bridged dimers, (6Clppdz)2Ir(μ-Cl)2Ir(6Clppdz)2 and (MCOMppdz)2Ir(μ-Cl)2Ir(MCOMppdz)2 in good yields. These chloride-bridged dimers are cleaved with 2-pyrazinecarboxylic acid (Hpyzc), 2-picolinic acid (Hpic), acetylacetone (Hacac), and 2,2,6,6-tetramethyl-3,5-heptanedione (Htmd) to give the corresponding monomer, (6Clppdz)2Ir(pyzc), (6Clppdz)2Ir(pic), (6Clppdz)2Ir(acac), (6Clppdz)2Ir(tmd), (MCOMppdz)2Ir(acac), and (MCOMppdz)2Ir(tmd), respectively. The complexes show bright orange to red luminescence at room temperature and the emission wavelenghts are affected by the ancillary ligands as well as cyclometalating ligands (593–664 nm).  相似文献   

11.
Zinc 2, (3)-tri-(phenylthio)-2, (3)-carboxy phthalocyanine (ZnPc(COOH)(SPh)3), zinc 2, (3)-tetra-(phenylthio) phthalocyanine (ZnPc(SPh)4) and 2, (3)-tetra-(phenylthio) phthalocyanine (H2Pc(SPh)4) were synthesized and their photophysical behavior were compared with those of a number of zinc phthalocyanine (ZnPc) derivatives. ZnPc(COOH)(SPh)3 and ZnPc(SPh)4 had similar fluorescence (ΦF=0.14) and triplet state (ΦT=0.65) quantum yields in dimethylsulfoxide, hence showing no effects of the replacement of one of the phenylthio groups with a carboxylic acid group. ZnPc(COOH)(SPh)3 displayed a slightly shorter triplet lifetime (τT=331 μs) than ZnPc (τT=350 μs) in DMSO, but within the range of ZnPc derivatives. The triplet lifetime for ZnPc(COOH)(SPh)3 is much longer than for the symmetrical derivative (ZnPc(SPh)4) with τT=149 μs in DMSO.  相似文献   

12.
The reductive and the oxidative electron-transfer photochemical reaction system of light-irradiated the mix solutions of Ru(bpy)3 2+ with [Fe(CN)6]4–, [Fe(CN)6]3–, [Fe(CN)5NO]2– and PB (Prussian Blue) have been studied. The double complexes which isolated from the precipitates of the photochemical reaction have been identified by means of Mössbauer spectroscopy. In order to clarify the chemical states of these isolated double complexes, we have (prepared and) studied Mössbauer spectra of the double complexes such as [Ru(bpy)3]3[Fe(CN)6]2.14H2O, [Ru(bpy)3]2[Fe(CN)6].10H2O, [Ru(bpy)3][Fe(CN)5NO].4H2O, and [Ru(bpy)3][PB]2.xH2O.  相似文献   

13.
ABSTRACT

A series of blue phosphorescent heteroleptic cyclometalated Ir(III) complexes with mesitylphenyl-imidazole ligands for organic light-emitting devices have been theoretically studied. We want to find their electronic structures, spectroscopic properties, and application value for organic light-emitting devices. (fppz)2Ir(acac), (fppz)2Ir(tpip), (dfbdp)2Ir(fppz), (F-fppz)2Ir(acac), (F-fppz)2Ir(tpip), and (dfbdp)2Ir(F-fppz) are investigated with DFT and TD-DFT approaches, where, for (fppz)2Ir(acac), (fppz denotes 2-(5-(trifluoromethyl)-4H-pyrazol-3-yl)pyridine, and acac denotes acetylacetonate); for (fppz)2Ir(tpip), tpip denotes tetraphenylimido-diphosphinate; and, for (F-fppz)2Ir(acac) and (F-fppz)2Ir(tpip), F-fppz denotes 2-(5-fluoro-4H-pyrazol-3-yl)pyridine.  相似文献   

14.
Nanostructured supramolecular complex of europium(III) with bathophenanthroline (bphen), with detonation synthesis nanodiamonds (NDs) used as a structure-forming element, has been synthesized for the first time. The characteristics of the Eu(bphen)2(NO3)3 complex and the supramolecular complex with NDs, ND–Eu(bphen)2(NO3)2, are studied and compared using scanning electron microscopy (SEM), luminescence spectroscopy, IR spectroscopy, and electron-spin resonance (ESR) spectroscopy. The luminescence quantum yields of the complexes are estimated by the relative method using a β-diketonate complex of europium (III) with tris(thenoyltrifluoroacetone) and 1,10-phenanthroline (Eu(TTA)3phen) as a reference. It is found that the ND–Eu(bphen)2(NO3)2 supramolecular complex has a higher photoresistance than the complex without NDs and no worse thermal stability (up to a temperature of +150°C).  相似文献   

15.
We analyze U(1)H as a horizontal symmetry and its possibilities to explain the known elementary-fermion masses. We find that only two candidates, in the context of SU(3)c ? SU(2)L ? U(1)Y ? U(1)H nonsuper-symmetric, are able to fit the experimental result mb«mt. identity, but it is a common prejudice to assume that the appropriate family symmetry may explain this fact as a consequence of (i) and (ii). In what follows we will enlarge the SM gauge group with an extra U(1)H horizontal local gauge symmetry (the simplest multi-family continuous symmetry we can think of). We then show that the structure SU(3)c ? SU(2)L ? U(1)Y ? U(1)H by itself is able to explain (ii), and that the simplest supersymmetric (SUSY) extension of this model without a μ-term can not cope with (ii).  相似文献   

16.
Potassium isopropyl xanthate, (CH3)2CHOC(S)SK, reacts with methyl chloroformiate ClC(O)OCH3 to yield (methoxycarbonyl) (2‐propoxythiocarbonyl) sulfide, (CH3)2CHOC(S)SC(O)OCH3. This novel xanthogen formate was characterized by 1H and 13C{1H} NMR spectroscopy, mass spectrometry and IR and Raman spectroscopy. The structure of a single crystal of (CH3)2CHOC(S)SC(O)OCH3 was determined by X‐ray diffraction analysis at 173 K. The conformational properties have been studied by liquid IR and Raman spectroscopy, matrix isolation spectroscopy together with photochemical studies and quantum chemical calculations (HF and B3LYP methods with the 6‐31+G* basis set). The analysis of the IR spectrum of liquid (CH3)2CHOC(S)SC(O)OCH3 suggests the presence of two conformers in equilibrium at room temperature. However, in the photochemical matrix study, an equilibrium of three conformers was detected. These forms were further characterized by theoretical calculations. Different photolysis products, such as CH3OC(O)SCH(CH3)2, OCS, CO, CO2 and CS2, were identified by matrix spectroscopy. The IR absorptions of CH3OC(O)SCH(CH3)2, for which literature data are scarce, were analysed in the light of the results of appropriate theoretical calculations. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

17.
The X-ray photoelectron spectra of M(N2)2 (dppe)2, M(N2)2 (PMePh2)4 and M(N2)2 (PMe2Ph)4; where M=W or Mo, Ph is phenyl, Me is methyl and dppe is (Ph)2P-CH2-CH2-P(Ph)2; have been investigated. The N(1s), P(2p3/2), Mo(3d5/2) and W(4f7/2) binding energies do not show significant differences between these complexes. The N(1s) signals show the separation of 1.3-1.4 eV.  相似文献   

18.
The thermal properties and the orientational order of hydrogen-bonded liquid crystals (6BA)2-(BPy)0.4 and (6BA)2-(BPy)0.3 (6BA: 4-n-hexylbenzoicacid, BPy: 4,4’-bipyridine) were investigated by DSC and 2H NMR. On cooling, isotropic liquid - liquid crystal phase transition temperatures were T C= 409 and 405 K for (6BA)2-(BPy)0.4 and (6BA)2-(BPy)0.3, respectively. Thermal anomalies in the liquid crystal phase were observed at T LC1= 402 and T LC2= 375 K for (6BA)2-(BPy)0.4 and at T LC1= 398 and T LC2= 375 K for (6BA)2-(BPy)0.3. For (6BA)2-(BPy)0.4, only the smectic component was created above T LC1. In addition, the nematic component was created below T LC1. The nematic component gradually changed to the smectic component with decreasing temperature and only the smectic component was observed below T LC2. For (6BA)2-(BPy)0.3, only the nematic component was created above T LC2. The phase transition from the nematic phase to the smectic phase took place at around T LC2.  相似文献   

19.
The structural and electronic properties of Cu5-1 and Cu6-1 nanowires with core-shell structures encapsulated inside a series zigzag (n,0) BeONTs denoted by Cu5-1@(n,0) and Cu6-1@(n,0) are investigated using the first-principles calculations within the generalized-gradient approximation. For Cu5-1@(n,0) (10 ? n ? 17) and Cu6-1@(n,0) (11 ? n ? 18) combined systems, the initial shapes (cylindrical BeONTs and CuNWs) are preserved without any visible change after optimization. The quantum conductances 5G 0 and 6G 0 of the most stable Cu5-1@(12,0) and Cu6-1@(13,0) combined systems are identical to the corresponding free-standing Cu5-1 and Cu6-1 nanowires, respectively. The energy bands crossing the Fermi level in both the Cu5-1@(12,0) and Cu6-1@(13,0) combined systems are all originated from the inner CuNWs. Therefore the electron transport will occur only through the inner CuNWs and the outer inert BeONTs serves well as an insulating cable sheath. The robust quantum conductance of the Cu5-1 and Cu6-1 nanowires, the insulating protection character of the (12,0) and (13,0) BeONTs and the highest stability of the tube-wire combined systems make the Cu5-1@(12,0) and Cu6-1@(13,0) combined systems are top-priority in the ULSI circuits and MEMS devices that demand steady transport of electrons.  相似文献   

20.
The microwave spectrum of the 35Cl and 37Cl isotopic species of 1-chloro-1,1,2-trifluoroethane (HCFC-133b) has been investigated in the frequency region 10 to 50 GHz using a Stark modulation microwave spectrometer. A pulsed jet Fourier transform microwave spectrometer was also used for the measurement of hyperfine splittings. A least-squares analysis of the observed b-type Q- and R-branch transition frequencies gave rotational and centrifugal distortion constants and components of the chlorine nuclear quadrupole coupling constant tensors in the principal axes system as follows: A=4625.161 (3) MHz, B=2004.127 (2) MHz, C=1875.813 (2) MHz, ΔJ=0.144 (9) kHz, ΔJK=1.0748 (8) kHz, ΔK=1.57 (1) kHz, δJ=0.01376 (4) kHz, δK=−0.146 (4) kHz, χaa=−57.958 (10) MHz, χbb=21.231 (11) MHz, and χcc=36.727 (11) MHz for 35ClCF2CH2F species, and A=4607.684 (6) MHz, B=1960.565 (2) MHz, C=1834.823 (2) MHz, ΔJ=0.106 (7) kHz, ΔJK=1.022 (3) kHz, ΔK=1.48 (1) kHz, δJ=0.0142 (2) kHz, δK=−0.18 (2) kHz, χaa=−46.268 (11) MHz, χbb=17.319 (13) MHz, and χcc=28.950 (13) MHz for 37ClCF2CH2F species. The structural parameters are calculated from the observed six rotational constants by assuming the partial structure of ab initio calculation. The electronic properties of the C-Cl bond are evaluated from the observed nuclear quadrupole constants of chlorine. These molecular properties are compared with those of other related molecules.  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