10.558=i 66.99=f 45...

若把数A的小数点向左移动一位后就得到数B,已知A+B=66.99,那么A=()。_百度知道
若把数A的小数点向左移动一位后就得到数B,已知A+B=66.99,那么A=()。
提问者采纳
已知A+B=66.99,那么A=( 60.9 )66.99÷ (10 +1)×10= 66.99 ÷11×10= 6.09×10= 60.9
提问者评价
其他类似问题
按默认排序
其他5条回答
若把数A的小数点向左移动一位后就得到数B,已知A+B=66.99,那么A=(60.9)。66.99÷(1+0.1)=60.9
A=66.99÷(10+1)=6.09
A=66。99/(1+0。1)=60。9
10A=B11A=66.99A=6.09B=60.9
小数点的相关知识
您可能关注的推广
等待您来回答
下载知道APP
随时随地咨询
出门在外也不愁手机查看产品信息
起订量:1 件
可售数量:面议
支持批发采购
山东省 聊城市
经营模式:
服务性企业
发送询价单,获取准确报价
分享拿好礼:
本企业产品分组
本企业推荐商品
同类商品推荐
国投华研研究院-过滤陶瓷片项目可行性研究报告
吸香水陶瓷片,陶瓷过滤片,吸油陶瓷棒(图)
吸香水陶瓷片,陶瓷过滤片,吸油陶瓷棒(图)
陶瓷过滤片,吸油陶瓷棒,吸香水陶瓷片,(图)
陶瓷过滤管 陶瓷过滤片 陶瓷过滤板 过滤陶瓷件
蜂窝陶瓷过滤片 微波烘干机
陶瓷垫片 陶瓷烟嘴 陶瓷过滤嘴 电子烟陶瓷座
蜂窝陶瓷过滤片
陶瓷垫片 陶瓷烟嘴 陶瓷过滤嘴 电子烟陶瓷座
特价供应金属过滤网片 优质空气净化过滤网 陶瓷过滤网 质优价廉
天津-矿用通信电缆MHYA32
丽水防辐射铅板生产厂家批发出厂价格
四川越西全国最大肉牛养牛基地
汕尾镇江隧道支护全自动混凝土喷浆机中铁局专用喷锚网
平安65mn钢切割零售厂家【】特约经销
没有找到想要的产品?
你是不是要找相关的产品信息?
发布询价单让供应商主动联系你
联系我时务必告知是在世界工厂网上看到的!
聊城市永红铸造工具有限公司
铸造用陶瓷过滤片的详细描述:
铸造用陶瓷过滤片
直孔陶瓷过滤片
产品说明;该产品具有比表面积大,机械强度高,耐热冲击的特点,通过型内过滤可取得以下效果:净化金属液,去除非金属杂质和气体;使金属液充型平稳,减少旋涡;简化浇注系统,提高工艺出品率;减少铸件气孔,优化细化金属组织;提高铸件表面质量和力学性能;降低铸件废品率;减少加工余量和提高刀具寿命,降低铸造成本。
挤出成型陶瓷滤渣片
&&&&—用于冶金、铸造行业金属熔融物过滤
&&&&该产品具有比表面积大,机械强度高,耐热冲击的特点,通过型内过滤可取得以下效果:净化金属液,去除非金属杂质和气体;使金属液充型平稳,减少旋涡;简化浇注系统,提高工艺出品率;减少铸件气孔,优化细化金属组织;提高铸件表面质量和力学性能;降低铸件废品率;减少加工余量和提高刀具寿命,降低铸造成本。
&&&&本公司在研究近年来国内、外铸造行业使用挤出型滤渣片所遇问题的基础上,有针对性地调整配方、创新工艺,生产出不同外型规格、不同壁厚、不同孔型和孔径的系列滤渣片产品。
堇青石、莫来石、锆莫来、锆刚玉不同材质过滤器最高使用温度分别为1350℃、1500℃、1580℃、1650℃
良好的抗热冲击性能,不会炸裂,不会被冲垮
均匀一致的圆形开孔使金属液通过时不会产生新的细小涡流,更利于实现完成金属液由紊流转变为层流的调整;
精度模具挤压成型带来的极强的产品一致性和质量稳定性。
同样外形规格的直通式陶瓷过滤器,方孔比圆孔开孔率高出20%左右
孔型状(round/square)
热膨胀系数(×10-6/℃)
软化温度(℃)
常温耐压强度(MPa)
刚玉莫来石
外形规格尺寸可由用户提供,也可共同设计、实验。
本公司现有过滤片几何特性
截面尺寸(长×宽mm)
孔数(个)
开孔率(%)
每立方片数
100×100×(20)
33×33=1089
100×100×(20)
31×31=961
100×100×(20)
24×24=576
81×81×(12.5)
1.46×1.46
44×44=1936
75×75×(15)
26×26=676
75×75×(15)
120×65×(20)
38×20=760
100×65×(20)
33×21=693
100×60×(12.5)
33×20=660
100×50×(20)
50×24=1200
66×66×(12.5)
42×42=1764
66×66×(12.5)
1.46×1.46
37×37=1369
66×66×(12.5)
26×26=676
66×66×(12.5)
60×60×(20)
29×29=841
60×60×(15)
20×20=400
55×55×(12.5)
55×55×(12.5)
21×21=441
55×55×(12.5)
1.15×1.15
37×37=1369
50×50×(12)
19×19=361
50×50×(12.5)
40×40×(10)
40×40×(10)
22×22=484
Φ90×(12.5)
Φ60×(10)
1.44×1.44
Φ52×(7)
Φ52×(7)
1.44×1.44
Φ50×(10)
温馨提示:
以上是关于铸造用陶瓷过滤片的详细介绍,
产品由聊城市永红铸造工具有限公司为您提供,如果您对聊城市永红铸造工具有限公司产品信息感兴趣可以
,您也可以查看更多与
相关的产品!
附件下载:
联系我时务必告知是在世界工厂网上看到的!
企业名称:
企业商铺:
联系地址:山东省聊城市-中国 山东 聊城 聊城市东昌府区 山东聊城
按字母分类:
相关区域产品:
免责声明:以上信息由会员自行提供,内容的真实性、准确性和合法性由发布会员负责,世界工厂网对此不承担任何责任。世界工厂网不涉及用户间因交易而产生的法律关系及法律纠纷,纠纷由您自行协商解决。
友情提醒:本网站仅作为用户寻找交易对象,就货物和服务的交易进行协商,以及获取各类与贸易相关的服务信息的平台。为避免产生购买风险,建议您在购买相关产品前务必确认供应商资质及产品质量。
联系方式:是处理侵权投诉的专用邮箱,在您的合法权益收到侵害时,欢迎您向该邮箱发送邮件,我们会在3个工作日内给您答复,感谢您对世界工厂网的关注与支持!Hindawi Publishing Corporation
Journal of Crystallography
Journal Menu
Journal of CrystallographyVolume ), Article ID
pagesResearch ArticleSynthesis, Characterization, and Crystal Structure of [Co4(CH3CO2)2L4]2[BPh4]4
 0.5H2O, Where HL = 4-(Salicylaldiminato)antipyrine,1 ,1 ,1 ,1 ,1 ,2 and 21Chemistry Department, Tripoli University, Tripoli,
Libya2Department of General and Inorganic Chemistry, Faculty of Technology and Metallurgy, University of Belgrade, Karnegijeva 4, 11000 Belgrade,
SerbiaReceived 15 April 2014; Accepted 21 June 2014; Published
27 August 2014Academic Editor: Yan Xu Copyright © 2014 Ramadan M. El-mehdawi
et al. This is an open access article distributed under the , which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.The title complex was isolated as a red solid from the reaction of 4-(salicylaldiminato)antipyrine, HL, and cobalt (II) acetate in ethanol. The complex has been characterized by elemental analysis, FTIR, UV-Vis, and X-ray single crystal diffraction. Two crystallographically different cationic units, A and B, of the title complex are found. Both units are essen nevertheless, small differences exist between them. Both units contain four cobalt atoms arranged at the corners of distorted cubane-like core alternatively with phenoxy oxygen of the Schiff base. In both cases, one cobalt binds to three coordinated sites from the corresponding tridentate Schiff base ligand, and the fourth one was bonded by the acetate oxygen, and the fifth and the sixth donor sites come from the phenolate oxygen of another Schiff base ligand.1. IntroductionPolynuclear as well as binuclear Co(II) complexes have attracted much attention due to their potential advantages than mononuclear complexes toward the preparation of molecular magnets and their application in data storage and memory devices []. The discovery of single molecule magnetism (SMM) in high-spin Ni(II) molecular clusters, particularly in cubane-like tetranuclear Ni(II) complexes, revived the interest in such compounds in order to study the correlation between the magnetic anisotropy of the high-spin ground state and the magnetization at low temperatures []. In comparison to mononuclear complexes, binuclear and polynuclear complexes could also provide more than one metal active center as Lewis acid in catalytic process, which is the interest of many researchers [].Crystal structures of complexes containing Co4O4 cubane-like core, as
 Co4O4(OAc)2(bpy)4
 (ClO4)2,
 Co4(CH3O)4(C5H7O2)4(C2H3O2)
 Co4O4(C8H9N2O2)
  7.5H2O [], are well-known for several decades. Polynuclear metal complexes with tridentate ligands containing at least one hydroxyl group and oxygen as terminal coordinating atom have been reported and attracted much attention []. These ligands often form polynuclear complexes with cubane or double cubane structure with missing one vertex each []. Antipyrine and its derivatives are one of such compounds that act as tridentate ligands, and this type of ligands has been attractive to researchers, since they are used as antifever and pain reliving drugs [].Transition metal Schiff base complexes have been also used as antifungal and antibacterial reagents []. It is shown that transition metal complexes containing 4-aminoantipyrine as a Schiff base have anticancer and antibacterial activity [] and have an important effect on simulate enzymes []. The type of the metal salt and the ligand plays an important role in tailoring the final product. Here, we report the first type of the Co(II) complex that contains two crystallographically different ions 1.a and 1.b, with tetranuclear cobalt atoms each. The Co(II) atoms reside in the corners of a cubane-like structure alternatively with oxygen atoms from the Schiff base ligands.2. Experimental2.1. Material and MeasurementAll reagent grade chemicals used in this work were obtained commercially from Aldrich or BDH and used without any further purification. All manipulations were carried out under atmospheric pressure. Elemental analysis (C, H, and N) was performed on a Vario El(III) elemental analyzer. FTIR spectra were recorded at room temperature with a Bruker IFS-25 OPUS/IR over the range from 400 to 4000 cm−1 with resolution of 4 cm−1. The electronic absorption spectrum was recorded over the range 200&#x nm using Cary 5000 UV-VISNR spectrophotometer.2.1.1. Synthesis of (4-Salicylaldiminato)antipyrine-EtOH (HL·EtOH)Yellow crystals of Schiff base, HL, were prepared using previously described method [, ]. A mixture of 4-aminoantipyrine (2.0 g, 9.85 mmol) in 20 mL ethanol and salicylaldehyde (1.22 g, 10 mmol) in 10 mL ethanol were refluxed together. Complete reaction with near quantitative conversion to the product required a period of 2.0 h. After cooling to room temperature, the yellow precipitate formed was collected by suction filtration and recrystallized from hot ethanol as deep yellow microcrystals (yield 2.81 g, 85%). The Schiff base was characterized by elemental analysis (Anal. Calc. for C20H22N3O3: C, 68.18; H, 6.25; N, 11.93; Found: C, 67.95; H, 6.25; N, 11.70).2.1.2. Synthesis of
 Co4(CH3CO2)2L4
 BPh4
 4·0.5H2OA solution of HL (2.0 mmol, 0.6 g) in 15 ml ethanol and cobalt (II) acetate (2.0 mmol) in 15 mL ethanol we after 2 h, a red solid material was precipitated. The resulting solid was separated by filtration, washed with ethanol and dichloromethane, and air-dried. The product was dissolved in hot methanol and a suitable methanol solution of sodium tetraphenylborate was added and stirred for a while to give the title complex as a red solid. The final product was dissolved in
on slow evaporation, red crystals were formed (yield 62%). The final product was characterized by elemental analysis, FTIR, UV-Vis, and single-crystal X-ray diffraction analysis (Anal. Calc. for C248H219B4Co8N24O24.5: C, 66.99; H, 4.93; N, 7.56. Found: C, 67.38; H, 5.4; N, 7.23).2.1.3. Crystal Structure Determination and Refinement of the Title ComplexSuitable red single crystals of the complex were obtained by slow evaporation of an acetone solution at room temperature. X-ray single-crystal diffraction data were collected on an Oxford Gemini S diffractometer equipped with CCD detector at 295 K; MoKα radiation (
  Å) was used and a multiscan correction for absorption was applied. The structure was solved by direct methods (SIR92) [] and refined on
  by full matrix least-squares SHELXL97 [] and WinGX []. All nonhydrogen atoms except
  were only partially occupied (s.o.f = 0.5) and their H atoms were not located. Positions of hydrogen atoms bonded to C were calculated geometrically and refined by a riding model. All aromatic C6-rings are constrained to be planar using FLAT instruction. Although there were indications that some rings are disordered, this behavior was not further explored due to very large number of atoms in the asymmetric unit. Data collection: CrysAlis PRO [], cell refinement: CrysAlis PRO; data reduction: CrysAlis PRO; program(s) used to solve structure: SIR97 [], program(s) used to refine structure: SHELXL97 [], molecular graphics: ORTEP-3 for Windows [], and Mercury [], software used to prepare material for publication: publCIF [], and PARST []. Crystal Data. The details of the crystallographic data and structure refinement for the complex are shown in Table . Additional material available from the Cambridge Crystallographic Data Center comprises thermal parameters and remaining bond distances and angles (CCDC number 801603).Table 1: Crystal data and structure refinement details for the complex.2.2. Crystal DataSee Table .3. Results and DiscussionA novel cobalt complex
 Co4(CH3CO2)2L4
 BPh4
 4·0.5H2O was synthesized by the reaction of cobalt (II) acetate with HL Schiff base in ethanolic solution. The IR spectrum of the complex exhibited strong absorptions at 1625 and 1562 cm−1 assignable to the carbonyl group of the pyrazolone ring, &#x3(C–O), and azomethine group (HC=N) of “L,” respectively. The first absorption band of the complex was shifted to the lower frequency for about 29 cm−1 relative to the free ligand (1654 cm−1) indicating that the ligand coordinates through the carbonyl oxygen of the pyrazolone ring. The second absorption was shifted to the lower frequency for about 29 cm−1 (free ligand at 1591 cm−1) suggesting the involvement of the nitrogen atom of the azomethine group in the coordination. The absorptions appears as shoulders at 1610 and 1457 cm−1 indicates the asymmetric and symmetric stretching frequencies of the acetate –
  groups. The separation (143 cm−1) between the two absorptions indicates that the acetate group acts as a bidentate bridging ligands []. The &#x3(OH) absorption at 3446 cm−1 indicates the presence of water molecule. The tetraphenylborate showed a finger print absorption at 2927&#x cm−1. The UV-Vis spectrum of the title complex was obtained in acetone solution. The spectrum shows a very intense band at 202 cm−1 due to
  transition characteristic for aromatic rings, and another intense band at 232 nm corresponding to the charge transfer band. The bands at 324 nm with a shoulder at 315 nm and 401 nm are assigned to the d-d transitions of the high-spin Co(II) ions in an octahedral geometry. There is a blue shift in this bands compared with the absorption spectra of some hexacoordinate Co(II) complexes with the peak maxima wavelength (
) at about 500&#x nm [].The asymmetric units of the title complex contain two tetranuclear cations, A and B, of the formula
 Co4(CH3CO2)2L4
 2+ with a cubane-like Co4O4 core. In addition, four
 BPh4
 − ions and one water molecule with 0.5 occupancy are present. The cations are close to be isostructural, and they only slightly differ in corresponding bond distances and angles. Crystal structure of the title complex with atomic numbering scheme are presented as ORTEP (Figure ), the stereoview of the cell contents of the title complex was shown in (Figure ) and relevant bond distances were given in Table . From ORTEP presentation and molecular packing diagram (Figure ), the ligand, “L,” acts as trichelate coordinating to the Co atom by two O and one N atoms. The oxygen of the phenolate group further bridges three Co(II) ions (&#x33-O) in the cubane-like core. Both cationic units contain Co(II) residing in a pseudooctahedral environment with CoNO5 chromophore. Each complex cation can be regarded as a dimer of dimeric units composed by phenoxy bridged dicobalt (II) subunits. The basal-planes of Co(II) include two oxygen atoms from the phenolate groups of two “L” ligands and another oxygen atom from the fourth position was occupied by the azomethine nitrogen of “L.” Axial sites are occupied by two oxygen atoms, one from the acetate group and the other one from another phenolate group of “L.”Table 2: Selected bond lengths (Å) and angles (°) of the title compound.Figure 1: An ORTEP drawing of
 Co4L4(OAc)2
  cationic units (a) and (b) of the title complex with 50% probability thermal ellipsoids showing the atom labeling scheme. Hydrogen atoms are omitted for clarity.Figure 2: Stereoview of the unit cell contents of
 Co4L4(OAc)2
 2(Bph4)4·0.5H2O. Hydrogen atoms are omitted for clarity.Figure 3: View of the molecular packing diagram showing 3D structure of the title complex. Hydrogen atoms are omitted for clarity.The acetate group acts as triatomic bidentate ligands bridging two Co atoms in their familiar syn-syn manner on opposite faces of the cube and the ligand “L” chelates the remaining four faces. The two faces bridged by acetate group in cationic unit A exhibit shorter Co≡Co separations of 2.996 and 3.022 Å, and more acute Co–O–Co angles ranging from 89.06 to 90.03° with Co–O–O–Co dihedral angles of 154.16 and 156.56° compared with those not bridged by acetate group, where Co≡Co separations are in the range of 3.24 to 3.31 Å, Co≡Co angles in the range of 97.26 to 102.28°, and Co–O–O–Co dihedral angles in the range of 177.06 to 178.99°. Cationic unit B is a little bit shorter in Co≡Co separations with more acute angles. These variations in bond distances and angles between faces of the cubane-like cores are responsible for their distortions. This distortion produces two types of Co–O–Co angles, those that are <90° at the top and bottom faces of the core and those that are >101° at the rest of the cubane faces. Therefore, one can expect that the Co(II) ions with small angles will lead to some magnetic properties comparable to the
 Ni4(OCH3)4(dbm)4(MeOH)4
, dbm = dibenzoylmethane [].The average Co–OAc bond distances of the title complex of 2.02 Å are longer than that reported for
 Co4O4(dpah)4(CH3CO2)2
 2V4O12·5H2O (1.98 Å) and shorter than that reported for
 Ni4(OCH3)4(OAc)2(TMB)4
 (BPh4)2·4CH2Cl2, TMP = 2,5-dimethyl-2,5-diisocyanohexane (2.05 Å) []. The bond distances of Co–OAc in cationic unit B are a little bit longer than in cationic unit A by 0.02 Å. The difference in all bond distances of the acetate groups in both units ranges from 0.01 to 0.02 Å and is very close to that reported for Ni(II) complex []. While the phenolate oxygen atom is ligating three cobalt atoms in a tetrahedral environment, the four cobalt atoms are located at four alternative corners of a distorted cubane-like core and the rest of the cubane vertices are occupied by four oxygen atoms from four L as &#x33-O. Both units (Figure ), as we mentioned before, differ slightly in corresponding bond
even in the cubane itself the separations between cobalt atoms differ slightly, for example, the longest separation for cationic unit A of Co2≡Co4 (3.288 Å) and the shortest separation of Co3≡Co4 (2.996 Å), for cationic unit B, the longest separation of Co6≡Co8 (3.313 Å), and the shortest of Co7≡Co8 (2.979 Å). The shortest Co≡Co separation in the title complex is still longer than the longest one reported for
 Co4O4(dpah)4(OAc)2
 2V4O12·5H2O, dpah = 2,2-dipyridylamine (2.848 Å) []. The average Co≡Co separations in both units are very close to the separation in
 Co2L2(NCS)2(CH3OH)2
  (3.1028 Å) [] and in
 Co2L2Cl2(CH5OH)2
  (3.085 Å) [], where “L” is the same Schiff base in the first and 2-hydroxyisophtaldehyde oxime in the second case. The C–O bond distances in the phenolate groups of the ligands, which act as &#x33-O bridges, range from 1.377(5) to 1.381(5) Å in cationic unit A and from 1.367(5) to 1.379(5) Å in cationic unit B. All of these distances are still longer than C–O distance of the phenolate group which acts as &#x32-O in
 Co2L2(NCS)2(CH3OH)2
  (1.330(2) Å) and even longer than the distance of the free Schiff base (1.345 Å) [, ].Further analysis of the crystal structure revealed that this structure contains a half water molecule in the lattice. There is a short contact between O25 and H27E atom belonging to the acetate methyl group at 3.01(1)8 Å and another O24
 O25 contact of 2.90(2) Å indicating possible hydrogen bonds. The four counter ions
 BPh4
 − exhibit their normal geometry. The two A and B cationic units of the title complex are held together by an electrostatic attraction with
 BPh4
 − ions. Finally, similar complexes of Ni(II), Mn(II), mixed Co(II) and Ni(II), and mixed Co(II) and Cu(II) complexes were prepared and characterized by spectroscopic methods and their structures will be investigated in near future.4. ConclusionFrom all of these, we conclude that the novel Co(II) complex has two very large A and B cationic unit complexes with a cubane- A and B units are almost isostructural and differ slightly in bond distance and angles. The Co(II) atoms reside in a pseudooctahedral environment and share the corners of a distorted cubane-like core equally with the oxygen of the phenolate group of the Schiff base ligands. The phenolate oxygen bridges three cobalt atoms, while the acetate group bridges two cobalt atoms. The distortion of the cationic unit complexes may lead to some magnetic interactions which is correlated with the small Co–O–Co angles of the top and bottom faces within the cubanes.Conflict of InterestsThe authors declare that there is no conflict of interests regarding the publication of this paper.
References
E. K. Brechin, C. Boskovic, W. Wernsdorfer et al., “Quantum tunneling of magnetization in a new [Mn18]2+ single-molecule magnet with S = 13,” Journal of the American Chemical Society, vol. 124, no. 33, pp. , 2002.
· M. Moragues-Cánovas, M. Helliwell, L. Ricard et al., “An Ni4 single-molecule magnet: synthesis, structure and low-temperature magnetic behavior,” European Journal of Inorganic Chemistry, vol. 2004, no. 11, pp. , 2004.
· J. Wang, J. Wu, and N. Tang, “Synthesis, characterization of a new bicobalt complex [Co2L2(C2H5OH)2Cl2] and application in cyclic carbonate synthesis,” Inorganic Chemistry Communications, vol. 10, no. 12, pp. , 2007.
· K. Dimitrov, K. Folting, W. E. Stereib, and G. Christo, “Dimerization of the [CoIII2(OH)2] core to the first example of a [CoIII4O4] cubane: potential insights into photosynthetic water oxidation,” Journal of the American Chemical Society, vol. 115, no. 14, pp. , 1993.
· M. Koikawa, H. Yamashita, and T. Tokii, “Synthesis, structure, and characterization of dicopper(II) complex with a new amidate ligand,” Inorganic Chemistry Communications, vol. 6, no. 2, pp. 157–161, 2003.
· A. E. Wasson and R. L. La Duca, “Hydrothermal synthesis, crystal structures, and properties of two 3-D network nickel nicotinate coordination polymers: [Ni4(&#x3-H2O)2(nicotinate)8· 2H2O] and [Ni2(H2O)2(nicotinate)4(4,4′-bpy)],” Polyhedron, vol. 26, no. 5, pp. , 2007.
· O. Y. Neiland, Organic Chemistry, Vysshaya Shkola, Moscow, Russia, 1990 (Russian). J. A. Ibers and R. H. Holm, “Modeling coordination sites in metallobiomolecules,” Science, vol. 209, no. 4453, pp. 223–235, 1980.
· M. T. H. Tarafder, M. A. Ali, D. J. Wee, et al., “Complexes of a tridentate ONS Schiff base. Synthesis and biological properties,” Transition Metal Chemistry, vol. 25, no. 4, pp. 456–460, 2000. A. Syamal and R. M. Maurya, “Oxidation of benzoin catalyzed by oxovanadium(iv) Schiff base,” Coordination Chemistry Reviews, vol. 18, p. 537, 1999. G. T. P. Britorsck, V. C. Gibson, and D. F. Wass, “The search for new-generation olefin polymerization catalysts,” Angewandte Chemie—International Edition, vol. 38, p. 428, 1999. Y. M. Chumakov, B. Y. Antosyak, M. D. Mazus, V. I. Taspkov, and M. N. Samus', “Crystal Structure of N-(salicylidene)-4-amino-2,3-dimethyl-1-phenyl-3-pyrazolin-5-one,” Journal of Structural Chemistry, vol. 41, no. 5, pp. 905–909, 2000.
· A. Altomare, G. Cascarano, C. Giacovazzo et al., “A program for automatic solution of crystal structures by direct methods,” Journal of Applied Crystallography, vol. 27, p. 435, 1994. G. M. Sheldrick, “A short history of SHELX,” Acta Crystallographica A, vol. 64, part 1, pp. 112–122, 2008.
· L. J. Farrugia, “WinGX suite for small-molecule single-crystal crystallography,” Journal of Applied Crystallography, vol. 32, pp. 837–838, 1999.
Oxford Diffraction, CrysAlis PRO, Version 1.171.33.31, Oxford Diffraction Ltd, England, UK, 2009. A. Altomare, M. C. Burla, M. Camalli et al., “SIR97: a new tool for crystal structure determination and refinement,” Journal of Applied Crystallography, vol. 32, no. 1, pp. 115–119, 1999.
· L. J. Farrugia, ORTEP-3 for Windows, University of Glasgow, Glasgow, Scotland, 1997. C. F. Macrae, P. R. Edgington, P. McCabe et al., “Mercury: visualization and analysis of crystal structures,” Journal of Applied Crystallography, vol. 39, no. 3, pp. 453–457, 2006.
· S. P. Westrip, “publCIF: software for editing, validating and formatting crystallographic information files,” Journal of Applied Crystallography, vol. 43, no. 4, pp. 920–925, 2010.
· M. Nardelli, “PARST95—an update to PARST: a system of Fortran routines for calculating molecular structure parameters from the results of crystal structure analyses,” Journal of Applied Crystallography, vol. 28, part 5, p. 659, 1995.
· R. H. Wang, D.-Q. Yuan, F.-L. Jiang, L. Han, S. Gao, and M. C. Hong, “Syntheses, structures, and characterization of two manganese(II)-aminobenzoic complexes,” European Journal of Inorganic Chemistry, vol. 2006, no. 8, pp. , 2006.
· L. Yaushiro, H. Nami, K. Hiroshi, and F. Shigenoba, “Electronic spectra of Co(II) Complexes,” Analytical Sciences, vol. 17, p. 187, 2001. M. A. Halerow, J.-S. Sun, J. C. Huffman, and G. Christou, “Structural and magnetic properties of [Ni4(.mu.3-OMe)4(dbm)4(MeOH)4] and [Ni4(.eta.1,.mu.3-N3)4(dbm)4(EtOH)4]. Magnetostructural correlations for [Ni4X4]4+ cubane complexes,” Inorganic Chemistry, vol. 34, no. 16, pp. , 1995.
· W. L. Gladfelter, M. W. Lynch, W. P. Schaefer, D. N. Henderikson, and H. B. Gray, “Synthesis, physical properties, and crystal structure of the cubane compound bis(.mu.-acetato)-tetra-.mu.-methoxo-bis[.mu.-(2,5-dimethyl-2,5-diisocyanohexane)]-tetranickel(II) tetraphenylborate,” Inorganic Chemistry, vol. 20, pp. , 1981.
· X. Zhang, W. You, Z. Zhu, L. Dang, Z. Sun, and X. Zheng, “Hydrothermal synthesis and characterization of a novel crystal containing [Co4O4]4+ cubane core: [Co4O4(dpaH)4(CH3COO)2]2V4O12· 5H2O,” Inorganic Chemistry Communications, vol. 9, no. 5, pp. 526–528, 2006.
· R. M. El-Mehdawi, N. A. Eldewik, K. M. Kreddan et al., “Synthesis and crystal structure of Bis-[Co (L)(NCS)(MeOH)] where (LH = 4-(Salicylaldiminato) antipyrine,” Jordan Journal of Chemistry, vol. 5, p. 157, 2010.

我要回帖

更多关于 186 7670 66.99 的文章

 

随机推荐