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 46 毫秒
1.
Two new one-dimensional Fe(II)-bis-Schiff base complexes, [Fe(L1)(pyz)] · CH2Cl2 (1) and [Fe(L2)(pyz)] · 2CH2Cl2 (2) (H2L1 = bis(O-vanillin)-O-phenylenediimine, H2L2 = bis(O-vanillin)-2,3-naphthalenediimine, pyz = pyrazine) are reported with their crystal structures and magnetic property. Compound 1 shows a two-step SCO behavior while 2 shows HS at all the temperature range measured. Although the extension of aromatic moiety from benzene (L1) to naphthalene (L2) was introduced for the purpose of strengthening the cooperativity, it leads to the absence of SCO, due to the unanticipated π–π interaction, which leads to the longer Fe–N bond lengths and a weak ligand field around Fe(II) ion.  相似文献   

2.
The synthesis and characterization of homobimetallic palladium and platinum complexes of type [(Me(O)CS-4-NCN–M  NN  M–NCN-4-SC(O)Me](OTf)2 (Me(O)CS-4-NCN = [C6H2(CH2NMe2)2-2,6-SC(O)Me-4]?; NN = 4,4′-bipyridine (bipy); M = Pd, 12; M = Pt, 13) is reported. The required bifunctional thio-acetyl NCN pincer starting compound NC(Br)N-4-SC(O)Me (2) has been synthesized by the consecutive reactions of NC(Br)N–I (I-1-C6H2(CH2NMe2)2-3,5-Br-4) (1) with tBuLi, S8 and Me(O)CCl, respectively. Chemoselective metallation at the Caryl–Br bond was achieved by the reaction of 2 with the palladium(0) source [Pd2(dba)3] (3) (dba = dibenzylidene acetone). Treatment of thus formed [Pd(NCN-4-SC(O)Me)(Br)] (4) with [AgOTf] (8) (OTf = triflate, OSO2CF3) gave [Pd(NCN-4-SC(O)Me)(H2O)][OTf] (9) which was further reacted with 0.5 equiv. of 4,4′-bipyridine (11a) to afford rigid-rod structured 12. When [Pt(tol)2(SEt2)]2 (5) (tol = 4-tolyl) was used instead of 3, then 13 was produced via the in situ formation of [PtBr(NCN-4-SC(O)Me)] (7) and [Pt(NCN-4-SC(O)Me)(H2O)][OTf] (10). Another possibility to synthesize 7 relied upon the subsequent reaction of 1 with 0.5 equiv. of 5 to give [PtBr(NCN-4-I)] (6) which further reacted with tBuLi, 1/8 S8 and Me(O)CCl to afford 7. The cyclic voltammograms of 2, 7, and 13 are discussed.Complex 7 was structurally characterized by single crystal X-ray crystallography. Organometallic 7 crystallizes with three independent molecules in the asymmetric unit and displays a monomeric structure as commonly encountered in d8-metal pincer chemistry.  相似文献   

3.
《Inorganica chimica acta》2006,359(7):2285-2290
Stopped-flow kinetic measurements were used to compare the reactivities of [Ru(medtra)(H2O)] (medtra3− = N-methylethylenediaminetriacetate) (1) and [Ru(hedtra)(H2O)] (2) (hedtra3− = N-hydroxyethylethylenediaminetriacetate) with NO in aqueous solution at 15 °C, pH 7.2 (phosphate buffer). The measured second-order rate constants (3 × 103 and 6 × 104 M−1 s−1 for 1 and 2, respectively) are three to four order of magnitudes lower than that for the reaction between [RuIII(edta)(H2O)] (3) with NO. However, NO scavenging studies of complexes 13, conducted by measuring the difference in nitrite production between treated and untreated murine macrophage cells, revealed that despite being less kinetically reactive toward NO, the [Ru(medtra)(H2O)] complex exhibited the highest NO scavenging ability and lowest toxicity of compounds 13.  相似文献   

4.
《Inorganica chimica acta》2006,359(5):1541-1548
The electronic structure of a series of 11 penta-coordinated dichloride mononuclear Mn(II) complexes [Mn(L)Cl2] (L = Cl-terpy, Br-terpy, OH-terpy, phenyl-terpy, tolyl-terpy, mesityl-terpy, EtO-terpy, Me2N-terpy, tBu3-terpy, py-phen, and dpya) has been investigated by a multifrequency EPR study (9–285 GHz). The X-ray structures of [Mn(Br-terpy)Cl2], [Mn(EtO-terpy)Cl2], [Mn(Me2N-terpy)Cl2] and [Mn(tolyl-terpy)Cl2] are described. The spin Hamiltonian parameters have been determined for all complexes and show that the steric and electronic effects of the N-tridentate ligand L do not induce appreciable variations on the zero field splitting parameters. The magnitude of D, close to 0.3 cm−1, is governed by the chloride anion. High-field EPR spectroscopy allows the determination of electronic parameters of mononuclear Mn(II) complexes characterized by relatively large magnitudes of D and the unambiguous interpretation of the X-band spectra of these kinds of complexes.  相似文献   

5.
The series of nitrosyl complexes trans-[Ru(NH3)4L(NO)]Cl3, L = caffeine, theophylline, imidazole and benzoimidazole in position trans to NO were prepared and their photochemical properties studied. All complexes showed nitric oxide (NO) release under light irradiation at 330–440 nm. Quantum yields for [Ru(NH3)4L(H2O)]3+ formation (?Ru(III)) were sensitive to the natures of L, λirr and pH. The major product of the irradiation of trans-[Ru(NH3)4L(NO+)]3+ is the trans-[RuIII(NH3)4L(Cl)]2+ and NO as suggested by UV–Vis, electrochemical, and FTIR techniques.  相似文献   

6.
Due to the similarity between Pd and Pt, the complexes of palladium(II) can be considered as potential anticancer agents. Activity of six PdCl2(X2Py)2 complexes (Py = pyridine, and X = CH3 or Cl) was measured by MTT test using MCF7, CCRF-SB, PC3 and human B-lymphoblastoid cell lines. We found that the effect of PdCl2(XnPy)2 was cell-specific and time-dependent. Obtained results were discussed and compared with the activation parameters calculated for the hydrolysis of PdCl2(XnPy)2, indicating a correlation between viability of MCF7 and CCRF-SB cells and rates of hydrolysis of the Pd(II) complexes.  相似文献   

7.
Reaction of the N-alkylaminopyrazole (NNN) ligands bis[(3,5-dimethyl-1-pyrazolyl)methyl]ethylamine (bdmae) and bis[(3,5-dimethyl-1-pyrazolyl)methyl]isopropylamine (bdmai) with [PdCl2(CH3CN)2] in a 1:1 M/L ratio in CH2Cl2 produces cis-[PdCl2(NNN)] (NNN = bdmae (1), bdmai (2)). The solid state structure of complex 1 was determined by X-ray diffraction studies. The bdmae ligand is coordinated through the two Npz atoms to the metal atom, which completes its coordination with two chlorine atoms in a cis disposition.Treatment of the corresponding ligand with [PdCl2(CH3CN)2] in 1:1 M/L ratio in the presence of AgBF4 and metathesis with NaBPh4 in CH2Cl2/CH3OH (3:1) gave [PdCl(bdmae)](BPh4) (3), and in the presence of NaBPh4 in CH2Cl2/CH3CN (3:1) gave [PdCl(bdmai)](BPh4) (4). Complexes 1 and 2 were again obtained when complexes 3 and 4 were heated under reflux in a solution of Et4NCl in acetonitrile. These Pd(II) compounds were characterised by elemental analyses, conductivity measurements, IR, 1H and 13C{1H} NMR, HMQC and NOESY spectroscopies. The NMR studies of the complexes prove the rigid conformation of the ligands when they are complexed.  相似文献   

8.
The linkage isomers [Re(NCS)6]2? and [Re(NCS)5(SCN)]2? are obtained by the reaction of [ReBr6]2? with NCS? in dimethylformamide. Some differences in the chemical behavior allowed their separation and structural characterization in the form of (NBu4)2[Re(NCS)6] (1) and [Zn(NO3)(Me2phen)2]2[Re(NCS)5(SCN)] (2), respectively (Bu = n-C4H9 and Me2phen = 2,9-dimethyl-1,10-phenanthroline).  相似文献   

9.
《Inorganica chimica acta》2006,359(5):1659-1665
The solvento calcium bromide CaBr2(DME) has been prepared by reacting anhydrous calcium acetate with acetyl bromide in dimethoxyethane (DME), while refluxing a suspension of CaCl2 in DME/SOCl2 gave CaCl2(DME), high yields being secured in both cases. Crystals of the ionic [CaBr(DME)2(H2O)2]Br (1), were shown to contain heptacoordinated calcium(II) with bidentate DME. Two polymorphs of the mononuclear, CaBr2(DME)2(MeCOOH) (2), differing for their packing efficiency, have been studied by X-ray diffractometry. Compound 2 loses one DME and the coordinated acetic acid under vacuum at 30 °C, yielding CaBr2(DME). Coordinated DME in CaX2(DME) was found to be substituted by nitrogen bases in toluene, the adducts CaX2py2 (X = Cl, Br), CaBr2(DMEDA)2, (DMEDA = N,N′-dimethylethylenediammine) and CaBr2(L–L–L) [L–L–L = 2,6-bis(3,5-dimethyl-N-pyrazolyl)pyridine)] being obtained in good yields.  相似文献   

10.
Four new palladium(II) complexes with the formula Pd(L)2, where L are quinoxaline-2-carbonitrile N1,N4-dioxide derivatives, were synthesized as a contribution to the chemistry and pharmacology of metal compounds with this class of pharmacologically interesting bioreductive prodrugs. Compounds were characterized by elemental, conductometric and thermogravimetric analyses, fast atom bombardment mass spectrometry (FAB-MS) and electronic, Fourier transform infrared (FTIR) and 1H-nuclear magnetic resonance spectroscopies. The complexes were subjected to cytotoxic evaluation on V79 cells in hypoxic and aerobic conditions. In addition, a preliminary study on interaction with plasmid DNA in normoxia was performed. Complexes showed different in vitro biological behavior depending on the nature of the substituent on the quinoxaline ring. Pd(L1)2 and Pd(L2)2, where L1 is 3-aminoquinoxaline-2-carbonitrile N1,N4-dioxide and L2 is 3-amino-6(7)-methylquinoxaline-2-carbonitrile N1,N4-dioxide, showed non selective cytotoxicity, being cytotoxic either in hypoxic or in aerobic conditions. On the other hand, Pd(L3)2, where L3 is 3-amino-6(7)-chloroquinoxaline-2-carbonitrile N1,N4-dioxide, resulted in vitro more potent cytotoxin in hypoxia (P = 5.0 μM) than the corresponding free ligand (P = 9.0 μM) and tirapazamine (P = 30.0 μM), the first bioreductive cytotoxic drug introduced into clinical trials. In addition, it showed a very good selective cytotoxicity in hypoxic conditions, being non-cytotoxic in normoxia. Its hypoxic cytotoxicity relationship value, HCR, was of the same order than those of other hypoxia selective cytotoxins (i.e., Mitomycine C, Misonidazole and the N-oxide RB90740). Interaction of the complexes with plasmid DNA in normoxia showed dose dependent ability to relax the negative supercoiled forms via different mechanisms. Pd(L2)2 introduced a scission event in supercoiled DNA yielding the circular relaxed form. Meanwhile, both Pd(L1)2 and Pd(L3)2 produced the loss of negative supercoils rendering a family of topoisomers with reduced electrophoretic mobility. Pd(L3)2 showed a more marked effect than Pd(L1)2. Indeed, for the highest doses assayed, Pd(L3)2 was even able to introduce positive supercoils on the plasmid DNA.  相似文献   

11.
《Inorganica chimica acta》2006,359(9):2835-2841
Rh(I) carbene complexes of [RhX(bmim)(η4-1,5-cod)] type (bmim = 1-butyl-3-methyl imidazolium cation, X = Cl 2, Br 3, I 4), obtained in the reaction of [Rh(OMe)(η4-1,5-cod)]2 (1) with [bmim]X ionic liquids, catalyzed polymerization of phenylacetylene (PA) to cis-polyphenylacetylene (PPA) in CH2Cl2 and in ionic liquids. The yield of PPA increased and molecular weight (Mw) decreased after addition of phosphorus ligands PPh3 or P(OPh)3. Complex 4 reacted with P(OPh)3 giving cis-[RhI(bmim)(P(OPh)3)2] (5) complex which catalyzed oligomerization but not polymerization of PA.  相似文献   

12.
《Inorganica chimica acta》2006,359(5):1650-1658
A series of nickel(II) and palladium(II) complexes containing one or two pentafluorophenyl ligands and the phosphino-amides o-Ph2PC6H4CONHR [R = iPr (a), Ph (b)] displaying different coordination modes have been synthesised. The chelating ability of these ligands and the influence of both coligands and the metal centre in their potential hemilabile behaviour have been explored. The crystal structure of (b) has been determined and reveals N–H⋯O intermolecular hydrogen bonding. Bis-pentafluorophenyl derivatives [M(C6F5)2(o-Ph2PC6H4CO-NHR)] [M = Ni; R = iPr (1a); R = Ph (1b); M = Pd; R = iPr (2a); R = Ph (2b)] in which (a) and (b) act as rigid P, O-chelating ligands were readily prepared from the labile precursors cis-[M(C6F5)2(PhCN)2]. X-ray structures of (1a), (1b) and (2a) have been established, allowing an interesting comparative structural discussion. Dinuclear [{Pd(C6F5)(tht)(μ-Cl)}2] reacted with (a) and (b) yielding the monopentafluorophenyl complexes [Pd(C6F5)Cl{PPh2(C6H4–CONH–R)}] (R = iPr (3a), Ph (3b)) that showed a P, O-chelating behaviour of the ligands, confirmed by the crystal structure determination of (3a). New cationic palladium(II) complexes in which (a) and (b) behave as P-monodentate ligands have been synthesised by reacting them with [{Pd(C6F5)(tht)(μ-Cl)}2], stoichiometric Ag(O3SCF3) and external chelating reagents such as cod [Pd(C6F5)(cod){PPh2(C6H4-CONH-R)}](O3SCF3)(R = iPr (4a), Ph (4b)) and 2,2-bipy [Pd(C6F5)(bipy){PPh2(C6H4-CONH-R)}](O3SCF3) (R = iPr (5a), Ph (5b)). When chloride abstraction in [{Pd(C6F5)(tht)(μ-Cl)}2] is promoted by means of a dithioanionic salt as dimethyl dithiophospate in the presence of (a) or (b), the corresponding neutral complexes [Pd(C6F5){S(S)P(OMe)2}{PPh2(C6H4-CONH-R)}] (R = iPr (6a), Ph (6b)) were obtained.  相似文献   

13.
《Inorganica chimica acta》2006,359(5):1351-1356
Energy-transfer rate-constants from photo-excited [Ru(N–N)3]2+ (N–N = 2,2′-bipyridine (bpy), 4,4′-dimethyl-2,2′-bipyridine (4dmb), 5,5′-dimethyl-2,2′-bipyridine (5dmb)) to [Cr(O–O)3]3− (O–O2− = ox2− ((COO)2), mal2− (CH2(COO)2)) and [Cr(CN)6]3− in encounter complexes were evaluated in aqueous solutions containing alkali metal ion. The rate constant depends on the molecular size of the ruthenium(II) complex: 1.8 × 108 s−1 for [Ru(bpy)3]2+ (molecular radius, r = 5.8 Å), 1.4 × 108 s−1 for [Ru(5dmb)3]2+ (r = 6.1 Å) and 0.96 × 108 s−1 for [Ru(4dmb)3]2+ (r = 6.7 Å) in the system of [Ru(N–N)3]2+–[Cr(ox)3]3− in aqueous solution. However, the rate constant is much more sensitive to the chromate(III) complex than to ruthenium(II) complex; 1.8 × 108 s−1 and 0.43 × 108 s−1 for [Cr(ox)3]3− (r = 4.0 Å) and [Cr(mal)3]3− (r = 4.2 Å) in the [Ru(bpy)3]2+–[Cr(O–O)3]3− systems, respectively. We conclude that the congeniality between the donor’s and acceptor’s ligands in encounter complex plays an important role in energy transfer in aqueous solution.  相似文献   

14.
《Inorganica chimica acta》2006,359(9):2896-2909
[RuCl3(NO)(P–P)], [P–P = R2P(CH2)nPR2 (n = 1–3) and R2P(CH2)POR2, PR2–CHCH–PR2, R = Ph and (C6H11)2P-(CH2)2-P(C6H11)2] were obtained and characterized by 31P {1H} NMR, IR spectroscopies and cyclic voltammetry. The structures of fac-[RuCl3(NO)(P–P)], P–P = dppm (1), dppe (2), c-dppen (3) and dppp (4), mer-[RuCl3(NO)(dcpe)] (6a) and mer-[RuCl3(NO)(dppmO)] (7) have been determined by X-ray diffraction. Photochemical isomerization of fac- to mer-[RuCl3(NO)(P–P)] was observed under white light in a CH2Cl2 solution and in solid state. The isomerization processes were followed by IR and 31P {1H} spectra. The mer-[RuCl3(15NO)(dppb)] isomer was used for the definition of the phosphorus atoms in the structure of the complex in solution. The electrochemical study shows that the oxidation/reduction processes observed in these complexes are dependent on both the isomer (fac or mer) and the solvent. In CH2Cl2, the NO+ reduction potentials are less negative for the mer-isomers than for the fac ones, while in CH3CN solvent these potentials are, in general, very close for both isomers.  相似文献   

15.
《Inorganica chimica acta》2006,359(6):1855-1869
A series of discrete, mononuclear palladium(II)–methyl complexes, together with several palladium(II)–chloro analogues, of pyridine-functionalised bis-NHC ligands have been prepared via ligand transmetallation from the silver(I)-NHC complexes. The reported complexes comprise examples with both the methylene-bridged 2,6-bis[(3-R-imidazolin-2-yliden-1-yl)methyl]pyridine (RCNC; R = Mes, dipp, tBu) and planar 2,6-bis(3-R-imidazolin-2-yliden-1-yl)pyridine (RCNC; R = Mes, dipp) ligands and, when combined with the previously reported MeCNC/MeCNC examples, cover a broad spectrum of ligand substituent steric and electronic properties, including the bulky Mes and dipp groups frequently used in catalytic applications. The palladium(II) complexes have been characterised by a variety of methods, including single crystal X-ray crystallography, with the shielding of the Pd–Me groups in the proton NMR spectra of some of the N-aryl substituted examples correlated with the proximity of the aryl rings to the methyl group in the solid state structures. The [PdMe(RCNC/RCNC)]+ complexes undergo thermal degradation via reductive methyl-NHC coupling to give 2-methyl-3-R-imidazolium-1-yl species with relative stabilities in the order of [PdMe(MesCNC)]BF4 > [PdMe(MeCNC)]BF4  [PdMe(MesCNC)]BF4 > [PdMe(MeCNC)]BF4 > [PdMe(tBuCNC)]BF4  [PdMe(tBuCNC)]BF4 (not isolable). A comparison of the activity of the complexes as precatalysts in a model Heck coupling reaction shows greatest activity in those species bearing bulkier N-substituents, with complexes bearing RCNC ligands generally more efficient precatalysts than those bearing RCNC ligands.  相似文献   

16.
A study of the electrochemical behavior of a series of antimetastatic mono- and di-ruthenium complexes, namely [Na][trans-RuIIICl4(DMSO)(L)] and [Na]2[{trans-RuIIICl4(DMSO)}2(μ-L)], L = pyrazine (pyz), pyrimidine (pym), 4,4′-bipyridine (bipy), and 1,2-bis-(4-pyridyl)ethylene (etbipy), is reported. The results obtained show that in all dimeric Ru(III) complexes linked by heterocyclic non-chelating N-donor bridges, the two redox centers behave independently (with no remarkable electrochemical interaction), thus conferring no advantage in the likely hypothesis they act as pro-drugs (activation by reduction). Moreover, electrochemical evaluation of interaction between albumin and the title complexes confirms that this protein can act as the vehicle for drugs of this type in blood.  相似文献   

17.
《Inorganica chimica acta》2006,359(11):3527-3534
The relative alkene dissociation energies and the structures of Pd(PH3)22-CH2CHX), trans-[Pd(PH3)2Cl(η2-CH2CHX)]+, trans-[Pd(PH3)Cl22-CH2CHX)], Cp2Zr(PH3)(η2-CH2CHX) and [Cp2Zr(CH3)(η2-CH2CHX)]+ (X = CN, Cl, Br, Me, OMe, NMe2) were calculated with the B3LYP density functional theory. We examined the correlations between the partial charges of the coordinated alkenes and the relative alkene dissociation energies. Through these correlations, we have been able to see how the alkene(π)-to-metal(d) donation and metal(d)-to-alkene(π*) back-donation interactions affect the relative alkene dissociation energies. We also examined the calculated structures and found that the Zr(IV) and Pd(II) complexes have a rather asymmetric alkene coordination while the Zr(II) and Pd(0) complexes have an approximately symmetric alkene coordination. The effects of the alkene(π)-to-metal(d) donation and metal(d)-to-alkene(π*) back-donation interactions on the structural features have also been discussed.  相似文献   

18.
Two dinuclear palladium(II) complexes, [{Pd(en)Cl}2(μ-pz)](NO3)2 and [{Pd(en)Cl}2(μ-pydz)](NO3)2, have been synthesized and characterized by elemental microanalysis and spectroscopic (1H and 13C NMR, IR and UV–vis) techniques (en is ethylenediamine; pz is pyrazine and pydz is pyridazine). The square planar geometry of palladium(II) metal centers in these complexes has been predicted by DFT calculations. The chlorido complexes were converted into the corresponding aqua complexes, [{Pd(en)(H2O)}2(μ-pz)]4+ and [{Pd(en)(H2O)}2(μ-pydz)]4+, and their reactions with N-acetylated l-histidylglycine (Ac–l–His–Gly) and l-methionylglycine (Ac–l–Met–Gly) were studied by 1H NMR spectroscopy. The palladium(II)-aqua complexes and dipeptides were reacted in 1:1 M ratio, and all reactions performed in the pH range 2.0 < pH < 2.5 in D2O solvent and at 37 °C. In the reactions of these complexes with Ac–l–His–Gly and Ac–l–Met–Gly dipeptides, the hydrolysis of the amide bonds involving the carboxylic group of both histidine and methionine amino acids occurs. The catalytic activities of the palladium(II)-aqua complexes were compared with those previously reported in the literature for the analogues platinum(II)-aqua complexes, [{Pt(en)(H2O)}2(μ-pz)]4+ and [{Pt(en)(H2O)}2(μ-pydz)]4+.  相似文献   

19.
The precursors bis[N-(alkyl)benzimidazoliumylmethyl]durene halide (1a: alkyl = C2H5, halide = Br?; 1b: alkyl = n-C4H9, halide = Cl?; durene = 1,2,4,5-tetramethylbenzene) and their two new NHC silver(I) complexes [Durene(CH2BimyEtAgBr)2] (2a) and [Durene(CH2BimynBuAgCl)2] (2b) (Bimy = benzimidazol-2-ylidene) have been prepared and characterized. In the crystal structures of 2a and 2b the aromatic π–π stacking interactions are observed.  相似文献   

20.
Two tetracyanometalate building blocks, [Fe(5,5′-dmbipy)(CN)4]? (2) and [Fe(4,4′-dmbipy)(CN)4]? (3) (5,5′-dmbipy = 5,5′-dimethyl-2,2′-bipyridine; 4,4′-dmbipy = 4,4′-dimethyl-2,2′-bipyridine), and two cyano-bridged heterobimetallic complexes, [Cu2(bpca)2(H2O)2Fe2(5,5′-dmbipy)2(CN)8] · 2[Cu(bpca)Fe(5,5′-dmbipy)(CN)4] · 4H2O (4) and [Cu(bpca)Fe(4,4′-dmbipy)(CN)4]n (5) (bpca = bis(2-pyridylcarbonyl)amidate), have been synthesized and structurally characterized. Complex 4 contains two dinuclear and one tetranuclear heterobimetallic clusters in an asymmetric unit whereas the structure of complex 5 features a one-dimensional heterobimetallic zigzag chain. The Cu(II) ion is penta-coordinated in the form of a distorted square-based pyramid. Magnetic studies show ferromagnetic coupling between Cu(II) and Fe(III) ions with g = 2.28, J1 = 2.64 cm?1, J2 = 5.40 cm?1 and TIP = ?2.36 × 10?3 for complex 4, and g = 2.17, J = 4.82 cm?1 and zJ = 0.029 cm?1 for complex 5.  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