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.
Phosphane and N‐heterocyclic carbene ligated gold(I) chlorides can be effectively activated by Na[Me3NB12Cl11] ( 1 ) under silver‐free conditions. This activation method with a weakly coordinating closo‐dodecaborate anion was shown to be suitable for a large variety of reactions known to be catalyzed by homogeneous gold species, ranging from carbocyclizations to heterocyclizations. Additionally, the capability of 1 in a previously unknown conversion of 5‐silyloxy‐1,6‐allenynes was demonstrated.  相似文献   

2.
Silver(I) salts of weakly coordinating anions (WCA) are commonly applied as oxidizing agents or halide abstracting reagents. The feasibility of a particular silver salt for such applications strongly depends on the “nakedness“ of the silver cation. In this study the reactivity of Ag[Me3NB12Cl11] in different solvents was investigated. Crystal structures of a variety of complexes were obtained. In several crystal structures two boron clusters are bridged by Ag–Cl contacts. This leads to polymeric structures (e.g. for Ag[Me3NB12Cl11]·0.5CH2Cl2 and Ag[Me3NB12Cl11]·SO2). Sterically demanding aromatics like mesitylene, pyrene, and acenaphthene are η1‐ or η2‐bonded to the silver atom and also form coordination polymers, whereas benzene as a ligand leads to a molecular structure, in which two benzene molecules are η2‐coordinated to the silver cation. In contrast, strong σ donor ligands like pyridine and triphenylphosphine give homoleptic silver complexes and thus cation and anion are separated. Furthermore, the ability of Ag[Me3NB12Cl11] for performing metathesis reactions was investigated. The reaction with AuICl gave the [Au(NCMe)2]+ cation.  相似文献   

3.
Carba‐closo‐dodecaborate anions with two functional groups have been synthesized via a simple two‐step procedure starting from monoamino‐functionalized {closo‐1‐CB11} clusters. Iodination at the antipodal boron atom provided access to [1‐H2N‐12‐I‐closo‐1‐CB11H10]? ( 1 a ) and [2‐H2N‐12‐I‐closo‐1‐CB11H10]? ( 2 a ), which have been transformed into the anions [1‐H2N‐12‐RC?C‐closo‐1‐CB11H10]? (R=H ( 1 b ), Ph ( 1 c ), Et3Si ( 1 d )) and [2‐H2N‐12‐RC?C‐closo‐1‐CB11H10]? (R=H ( 2 b ), Ph ( 2 c ), Et3Si ( 2 d )) by microwave‐assisted Kumada‐type cross‐coupling reactions. The syntheses of the inner salts 1‐Me3N‐12‐RC?C‐closo‐1‐CB11H10 (R=H ( 1 e ), Et3Si ( 1 f )) and 2‐Me3N‐12‐RC?C‐closo‐1‐CB11H10 (R=H ( 2 e ), Et3Si ( 2 f )) are the first examples for a further derivatization of the new anions. All {closo‐1‐CB11} clusters have been characterized by multinuclear NMR and vibrational spectroscopy as well as by mass spectrometry. The crystal structures of Cs 1 a , [Et4N] 2 a , K 1 b , [Et4N] 1 c , [Et4N] 2 c , 1 e , and [Et4N][1‐H2N‐2‐F‐12‐I‐closo‐1‐CB11H9]?0.5 H2O ([Et4N ]4 a ?0.5 H2O) have been determined. Experimental spectroscopic data and especially spectroscopic data and bond properties derived from DFT calculations provide some information on the importance of inductive and resonance‐type effects for the transfer of electronic effects through the {closo‐1‐CB11} cage.  相似文献   

4.
The closo‐dodecaborate [B12H12]2? is degraded at room temperature by oxygen in an acidic aqueous solution in the course of several weeks to give B(OH)3. The degradation is induced by Ag2+ ions, generated from Ag+ by the action of H2S2O8. Oxa‐nido‐dodecaborate(1?) is an intermediate anion, that can be separated from the reaction mixture as [NBzlEt3][OB11H12] after five days in a yield of 18 %. The action of FeCl3 on the closo‐undecaborate [B11H11]2? in an aqueous solution gives either [B22H22]2? (by fusion) or nido‐B11H13(OH)? (by protonation and hydration), depending on the concentration of FeCl3. In acetonitrile, however, [B11H11]2? is transformed into [OB11H12]? by Fe3+ and oxygen. The radical anions [B12H12] ˙ ? and [B11H11] ˙ ? are assumed to be the primary products of the oxidation with the one‐electron oxidants Ag2+ and Fe3+, respectively. These radical anions are subsequently transformed into [OB11H12]? by oxygen. The crystal structure analysis shows that the structure of [OB11H12]? is derived from the hypothetical closo‐oxaborane OB12H12 by removal of the B3 vertex, leaving a non‐planar pentagonal aperture with a three‐coordinate O vertex, as predicted by NMR spectra and theory.  相似文献   

5.
Ten [C8C1Im]+ (1‐methyl‐3‐octylimidazolium)‐based ionic liquids with anions Cl?, Br?, I?, [NO3]?, [BF4]?, [TfO]?, [PF6]?, [Tf2N]?, [Pf2N]?, and [FAP]? (TfO=trifluoromethylsulfonate, Tf2N=bis(trifluoromethylsulfonyl)imide, Pf2N=bis(pentafluoroethylsulfonyl)imide, FAP=tris(pentafluoroethyl)trifluorophosphate) and two [C8C1C1Im]+ (1,2‐dimethyl‐3‐octylimidazolium)‐based ionic liquids with anions Br? and [Tf2N]? were investigated by using X‐ray photoelectron spectroscopy (XPS), NMR spectroscopy and theoretical calculations. While 1H NMR spectroscopy is found to probe very specifically the strongest hydrogen‐bond interaction between the hydrogen attached to the C2 position and the anion, a comparative XPS study provides first direct experimental evidence for cation–anion charge‐transfer phenomena in ionic liquids as a function of the ionic liquid’s anion. These charge‐transfer effects are found to be surprisingly similar for [C8C1Im]+ and [C8C1C1Im]+ salts of the same anion, which in combination with theoretical calculations leads to the conclusion that hydrogen bonding and charge transfer occur independently from each other, but are both more pronounced for small and more strongly coordinating anions, and are greatly reduced in the case of large and weakly coordinating anions.  相似文献   

6.
The deprotonation of the nido‐anion [B11H14] by two equivalents of LitBu yields the anion [B11H12]3–. Three observed 11B NMR shifts of this anion in the ratio 1 : 5 : 5 are in agreement with shifts calculated by the GIAO method on the basis of the ab initio computed geometry. The deprotonation can be reversed, giving back [B11H14] via [B11H13]2–. The thermolysis of [Li(thp)x]3[B11H12] in thp at 80 °C leads to the closo‐borate [Li(thp)3]2[B11H11] under elimination of LiH. Anhydrous air transforms [B11H12]3– into the known oxa‐nido‐dodecaborate [OB11H12]. The rhoda‐closo‐dodecaborate [L2RhB11H11]3– is formed from [B11H12]3– and RhL3Cl (L = PPh3).  相似文献   

7.
The hydride-bridged silylium cation [Et3Si−H−SiEt3]+, stabilized by the weakly coordinating [Me3NB12Cl11] anion, undergoes, in the presence of excess silane, a series of unexpected consecutive reactions with the valence-isoelectronic molecules CS2 and CO2. The final products of the reaction with CS2 are methane and the previously unknown [(Et3Si)3S]+ cation. To gain insight into the entire reaction cascade, numerous experiments with varying conditions were performed, intermediate products were intercepted, and their structures were determined by X-ray crystallography. Besides the [(Et3Si)3S]+ cation as the final product, crystal structures of [(Et3Si)2SMe]+, [Et3SiS(H)Me]+, and [Et3SiOC(H)OSiEt3]+ were obtained. Experimental results combined with supporting quantum-chemical calculations in the gas phase and solution allow a detailed understanding of the reaction cascade.  相似文献   

8.
The azaborate K2[nido-NB10H11] is gained from nido-NB10H13 and K[BHEt3] in a 1:2 ratio. The anion [NB10H11]2?, which is isoelectronic with [C2B9H11]2?, reacts with [{η6-(C6R6) · RuCl2}2] (R = H, Me), [{η5-(C5Me5)RhCl2}2], or [Ni(PPh3)2Cl2] to give the azametalla-closo-dodecaboranes MNB10H11 with M = (C6Me6)Ru ( 2 ), (C6H6)Ru ( 3 ), (C5Me5)Rh ( 4 ), and (Ph3P)2Ni ( 5 ), respectively. The azametallaborane K[Co(NB10H11)2] ( 6 ), which contains a sandwich-type coordinated Co atom, is formed from K2[NB10H11] and CoCl2. The structure of 2 · CH2Cl2 was determined by X-ray diffraction. The products 2 – 6 can be derived from the icosahedral anion [B12H12]2? on replacing a BH2? moiety by the isoelectronic nitrene NH and a BH moiety by the isolobal metal-complex fragment M. The N atom is six-coordinated in the cluster skeletons 2 – 6 .  相似文献   

9.
The anionic gold(I) complexes [1‐(Ph3PAu)‐closo‐1‐CB11H11]? ( 1 ), [1‐(Ph3PAu)‐closo‐1‐CB9H9]? ( 2 ), and [2‐(Ph3PAu)‐closo‐2‐CB9H9]? ( 3 ) with gold–carbon 2c–2e σ bonds have been prepared from [AuCl(PPh3)] and the respective carba‐closo‐borate dianion. The anions have been isolated as their Cs+ salts and the corresponding [Et4N]+ salts were obtained by salt metathesis reactions. The salt Cs‐ 3 isomerizes in the solid state and in solution at elevated temperatures to Cs‐ 2 with ΔHiso=(?75±5) kJ mol?1 (solid state) and ΔH=(118±10) kJ mol?1 (solution). The compounds were characterized by vibrational and multi‐NMR spectroscopies, mass spectrometry, elemental analysis, and differential scanning calorimetry. The crystal structures of [Et4N]‐ 1 , [Et4N]‐ 2 , and [Et4N]‐ 3 were determined. The bonding parameters, NMR chemical shifts, and the isomerization enthalpy of Cs‐ 3 to Cs‐ 2 are compared to theoretical data.  相似文献   

10.
A simple method for the functionalization of closo‐borates [closo‐B10H10]2? ( 1 ), [closo‐1‐CB9H10]? ( 2 ), [closo‐B12H12]2? ( 3 ), [closo‐1‐CB11H12]? ( 4 ), and [3,3′‐Co(1,2‐C2B9H11)2]? ( 5 ) is described. Treatment of the anions and their derivatives with ArI(OAc)2 gave aryliodonium zwitterions, which were sufficiently stable for chromatographic purification. The reactions of these zwitterions with nucleophiles provided facile access to pyridinium, sulfonium, thiol, carbonitrile, acetoxy, and amino derivatives. The synthetic results are augmented by mechanistic considerations.  相似文献   

11.
Reactions of bis(phosphinimino)amines LH and L′H with Me2S ? BH2Cl afforded chloroborane complexes LBHCl ( 1 ) and L′BHCl ( 2 ), and the reaction of L′H with BH3 ? Me2S gave a dihydridoborane complex L′BH2 ( 3 ) (LH=[{(2,4,6‐Me3C6H2N)P(Ph2)}2N]H and L′H=[{(2,6‐iPr2C6H3N)P(Ph2)}2N]H). Furthermore, abstraction of a hydride ion from L′BH2 ( 3 ) and LBH2 ( 4 ) mediated by Lewis acid B(C6F5)3 or the weakly coordinating ion pair [Ph3C][B(C6F5)4] smoothly yielded a series of borenium hydride cations: [L′BH]+[HB(C6F5)3]? ( 5 ), [L′BH]+[B(C6F5)4]? ( 6 ), [LBH]+[HB(C6F5)3]? ( 7 ), and [LBH]+[B(C6F5)4]? ( 8 ). Synthesis of a chloroborenium species [LBCl]+[BCl4]? ( 9 ) without involvement of a weakly coordinating anion was also demonstrated from a reaction of LBH2 ( 4 ) with three equivalents of BCl3. It is clear from this study that the sterically bulky strong donor bis(phosphinimino)amide ligand plays a crucial role in facilitating the synthesis and stabilization of these three‐coordinated cationic species of boron. Therefore, the present synthetic approach is not dependent on the requirement of weakly coordinating anions; even simple BCl4? can act as a counteranion with borenium cations. The high Lewis acidity of the boron atom in complex 8 enables the formation of an adduct with 4‐dimethylaminopyridine (DMAP), [LBH ? (DMAP)]+[B(C6F5)4]? ( 10 ). The solid‐state structures of complexes 1 , 5 , and 9 were investigated by means of single‐crystal X‐ray structural analysis.  相似文献   

12.
Most of the divalent compounds of tin have a lone pair and hence can act as donors. In tin‐transition metal chemistry neutral molecules as well as anions have been studied as ligands. This research report summarizes recent research on coordination compounds with a closo‐heteroborate cage compound stanna‐closo‐dodecaborate [SnB11H11]2?. The syntheses of the first coordination compounds and studies on the ligand abilities of this tin borate are discussed in this article.  相似文献   

13.
Reaction of CsF with ClF3 leads to Cs[Cl3F10]. It contains a molecular, propeller‐shaped [Cl3F10]? anion with a central μ3‐F atom and three T‐shaped ClF3 molecules coordinated to it. This anion represents the first example of a heteropolyhalide anion of higher ClF3 content than [ClF4]? and is the first Cl‐containing interhalogen species with a μ‐bridging F atom. The chemical bonds to the central μ3‐F atom are highly ionic and quite weak as the bond lengths within the coordinating XF3 units (X = Cl, and also calculated for Br, I) are almost unchanged in comparison to free XF3 molecules. Cs[Cl3F10] crystallizes in a very rarely observed A[5]B[5] structure type, where cations and anions are each pseudohexagonally close packed, and reside, each with coordination number five, in the trigonal bipyramidal voids of the other.  相似文献   

14.
The unexpected but facile preparation of the silver salt of the least coordinating [(RO)3Al‐F‐Al(OR)3]? anion (R=C(CF3)3) by reaction of Ag[Al(OR)4] with one equivalent of PCl3 is described. The mechanism of the formation of Ag[(RO)3Al‐F‐Al(OR)3] is explained based on the available experimental data as well as on quantum chemical calculations with the inclusion of entropy and COSMO solvation enthalpies. The crystal structures of (RO)3Al←OC4H8, Cs+[(RO)2(Me)Al‐F‐Al(Me)(OR)2]?, Ag(CH2Cl2)3+[(RO)3Al‐F‐Al(OR)3]? and Ag(η2‐P4)2+[(RO)3Al‐F‐Al(OR)3]? are described. From the collected data it will be shown that the [(RO)3Al‐F‐Al(OR)3]? anion is the least coordinating anion currently known. With respect to the fluoride ion affinity of two parent Lewis acids Al(OR)3 of 685 kJ mol?1, the ligand affinity (441 kJ mol?1), the proton and copper decomposition reactions (?983 and ?297 kJ mol?1) as well as HOMO level and HOMO–LUMO gap and in comparison with [Sb4F21]?, [Sb(OTeF5)6]?, [Al(OR)4]? as well as [B(RF)4]? (RF=CF3 or C6F5) the [(RO)3Al‐F‐Al(OR)3]? anion is among the best weakly coordinating anions (WCAs) according to each value. In contrast to most of the other cited anions, the [(RO)3Al‐F‐Al(OR)3] anion is available by a simple preparation in conventional inorganic laboratories. The least coordinating character of this anion was employed to clarify the question of the ground state geometry of the Ag(η2‐P4)2+ cation (D2h, D2 or D2d?). In agreement with computational data and NMR spectra it could be shown that the rotation along the Ag‐(P‐P‐centroid) vector has no barrier and that the structure adopted in the solid state depends on packing effects which lead to an almost D2h symmetric Ag(η2‐P4)2+ cation (0 to 10.6° torsion) for the more symmetrical [Al(OR)4]? anion, but to a D2 symmetric Ag(η2‐P4)2+ cation with a 44° twist angle of the two AgP2 planes for the less symmetrical [(RO)3Al‐F‐Al(OR)3]? anion. This implies that silver back bonding, suggested by quantum chemical population analyses to be of importance, is only weak.  相似文献   

15.
A series of gold acetonitrile complexes [Au(NCMe)2]+[WCA]? with weakly coordinating counterions (WCAs) was synthesized by the reaction of elemental gold and nitrosyl salts [NO]+[WCA]? in acetonitrile ([WCA]? = [GaCl4]?, [B(CF3)4]?, [Al(ORF)4]?; RF = C(CF3)3). In the crystal structures, the [Au(NCMe)2]+ units appeared as monomers, dimers, or chains. A clear correlation between the aurophilicity and the coordinating ability of counterions was observed, with more strongly coordinating WCAs leading to stronger aurophilic contacts (distances, C?N stretching frequencies of [Au(NCMe)2]+ units). An attempt to prepare [Au(L)2]+ units, even with less weakly basic solvents like CH2Cl2, led to decomposition of the [Al(ORF)4]? anion and formation of [NO(CH2Cl2)2]+[F(Al(ORF)3)2]?. All nitrosyl reagents [NO]+[WCA]? were generated according to an optimized procedure and were thoroughly characterized by Raman and NMR spectroscopy. Moreover, the to date unknown species [NO]+[B(CF3)3CN]? was prepared. Its reaction with gold unexpectedly produced [Au(NCMe)2]+[Au(NCB(CF3)3)2]?, in which the cyanoborate counterion acts as an anionic ligand itself. Interestingly, the auroborate anion [Au(NCB(CF3)3)2]? behaves as a weakly coordinating counterion, which becomes evident from the crystallographic data and the vibrational spectral characteristics of the [Au(NCMe)2]+ cation in this complex. Ligand exchange in the only room temperature stable salt of this series, [Au(NCMe)2]+[Al(ORF)4]?, is facile and, for example, [Au(PPh3)(NCMe)]+[Al(ORF)4]? can be selectively generated. This reactivity opens the possibility to generate various [AuL1L2]+[Al(ORF)4]? salts through consecutive ligand‐exchange reactions that offer access to a huge variety of AuI complexes for gold catalysis.  相似文献   

16.
The Lanthanum Dodecahydro‐closo‐Dodecaborate Hydrate [La(H2O)9]2[B12H12]3·15 H2O and its Oxonium‐Chloride Derivative [La(H2O)9](H3O)Cl2[B12H12]·H2O By neutralization of an aqueous solution of the free acid (H3O)2[B12H12] with basic La2O3 and after isothermic evaporation colourless, face‐rich single crystals of a water‐rich lanthanum(III) dodecahydro‐closo‐dodecaborate hydrate [La(H2O)9]2[B12H12]3·15 H2O are isolated. The compound crystallizes in the trigonal system with the centrosymmetric space group (a = 1189.95(2), c = 7313.27(9) pm, c/a = 6.146; Z = 6; measuring temperature: 100 K). The crystal structure of [La(H2O)9]2[B12H12]3·15 H2O can be characterized by two of each other independent, one into another posed motives of lattice components. The [B12H12]2− anions (d(B–B) = 177–179 pm; d(B–H) = 105–116 pm) are arranged according to the samarium structure, while the La3+ cations are arranged according to the copper structure. The lanthanum cations are coordinated in first sphere by nine oxygen atoms from water molecules in form of a threecapped trigonal prism (d(La–O) = 251–262 pm). A coordinative influence of the [B12H12]2− anions on La3+ has not been determined. Since “zeolitic” water of hydratation is also present, obviously the classical H–Oδ–···H–O‐hydrogen bonds play a significant role in the stabilization of the crystal structure. During the conversion of an aqueous solution of (H3O)2[B12H12] with lanthanum trichloride an anion‐mixed salt with the composition [La(H2O)9](H3O)Cl2[B12H12]·H2O is obtained. The compound crystallizes in the hexagonal system with the non‐centrosymmetric space group (a = 808.84(3), c = 2064.51(8) pm, c/a = 2.552; Z = 2; measuring temperature: 293 K). The crystal structure can be characterized as a layer‐like structure, in which [B12H12]2− anions and H3O+ cations alternate with layers of [La(H2O)9]3+ cations (d(La–O) = 252–260 pm) and Cl anions along [001]. The [B12H12]2− (d(B–B) = 176–179 pm; d(B–H) = 104–113 pm) and Cl anions exhibit no coordinative influence on La3+. Hydrogen bonds are formed between the H3O+ cations and [B12H12]2− anions, also between the water molecules of [La(H2O)9]3+ and Cl anions, which contribute to the stabilization of the crystal structure.  相似文献   

17.
The first primary 2‐aminocarba‐closo‐dodecaborates [1‐R‐2‐H2N‐closo‐CB11H10]? (R=H ( 1 ), Ph ( 2 )) have been synthesized by insertion reactions of (Me3Si)2NBCl2 into the trianions [7‐R‐7‐nido‐CB10H10]3?. The difunctionalized species [1,2‐(H2N)2closo‐CB11H10] ( 3 ) and 1‐CyHN‐2‐H3N‐closo‐CB11H10 (H‐ 4 ) have been prepared analogously from (Me3Si)2NBCl2 and 7‐H3N‐7‐nido‐CB10H12. In addition, the preparation of [Et4N][1‐H2N‐2‐Ph‐closo‐CB11H10] ([Et4N]‐ 5 ) starting from PhBCl2 and 7‐H3N‐7‐nido‐CB10H12 is described. Methylation of the [1‐Ph‐2‐H2N‐closo‐CB11H10]? ion ( 2 ) to produce 1‐Ph‐2‐Me3N‐closo‐CB11H10 ( 6 ) is reported. The crystal structures of [Et4N]‐ 2 , [Et4N]‐ 5 , and 6 were determined and the geometric parameters were compared to theoretical values derived from DFT and ab initio calculations. All new compounds were studied by NMR, IR, and Raman spectroscopy, MALDI mass spectrometry, and elemental analysis. The discussion of the experimental NMR chemical shifts and of selected vibrational band positions is supported by theoretical data. The thermal properties were investigated by differential scanning calorimetry (DSC). The pKa values of 2‐H3N‐closo‐CB11H11 (H‐ 1 ), 1‐H3N‐closo‐CB11H10 (H‐ 7 ), and 1,2‐(H3N)2closo‐CB11H10 (H2‐ 3 ) were determined by potentiometric titration and by NMR studies. The experimental results are compared to theoretical data (DFT and ab initio). The basicities of the aminocarba‐closo‐dodecaborates agree well with the spectroscopic and structural properties.  相似文献   

18.
The lithium salt of the weakly coordinating alkoxyaluminate anion Li[Al(OC(CF3)2(CH2SiMe3))4] ( 2 ) is soluble in polar and even in non‐polar solvents. Especially the solubility in n‐hexane confirms 2 to be an excellent candidate for Li ion catalysis. Its polymeric structure consists of a seven coordinated Li+ cation, coordinating a [Al(OC(CF3)2(CH2SiMe3)]? anion that serves as hexadentate O2F4 ligand and a further bridging F atom of a second anion. Compound 2 reacts with ClCPh3 giving the [CPh3]+ salt which is at least stable in CD2Cl2 over days at 298 K, but decomposes after storage at 333 K for several days.  相似文献   

19.
Two novel azido‐derivatives of closo‐dodecaborate anion with hydrophobic and hydrophilic spacers were prepared by reaction of tetrabutylammonium azide with cyclic oxonium derivatives of the closo‐dodecaborate anion. The compounds prepared can be regarded as precursors of derivatives of closo‐dodecaborate anion with amino group at the terminal position of a spacer or as building blocks for ‘click chemistry’, which are useful for preparation of various conjugates with targeting molecules. A concentration dependence of the 11B NMR spectra of functionalized derivatives of closo‐dodecaborate anion was discovered, which is of great importance for analytical purposes. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

20.
The lithiacarborane [Li?CB11H11]? plays a central role in carborane chemistry, as it is a key intermediate to achieve the selective functionalization of the monocarba‐closo‐dodecaborate [CB11H12]? for applications in various fields. Also, it is an organometallic species of fundamental interest because it represents a 3D analogue of phenyllithium featuring an exo C?Li bond in addition to the delocalized negative endo charge of the spherical cluster. For the first time, the elusive and highly reactive endo/exo formal dianion [CB11H11]2? has been isolated as its lithiate as well as zincate in pure form and fully characterized. DFT calculations corroborate the experimental findings and underscore the remarkably high reactivity of the lithiacarborane. Subsequent derivatizations demonstrate the relevance of its initial clean formation.  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