Journal articles on the topic 'X-ray crystallographically'

To see the other types of publications on this topic, follow the link: X-ray crystallographically.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'X-ray crystallographically.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Aitken, R. Alan, Fiona M. Fotherby, and Alexandra M. Z. Slawin. "2,6-exo-8,12-exo-10-Butyl-13-oxa-3,5-dithia-10-azatetracyclo[5.5.1.02,6.08,12]tridecane-9,11-dione." Molbank 2020, no. 2 (April 2, 2020): M1123. http://dx.doi.org/10.3390/m1123.

Full text
Abstract:
The title compound was obtained in low yield and spectroscopically characterised. Its X-ray structure was compared with the X-ray structures of other crystallographically-characterised 2-unsubstituted 1,3-dithiolanes.
APA, Harvard, Vancouver, ISO, and other styles
2

Anzaldo, Bertin, Pankaj Sharma, Francisco Lara Ochoa, Claudia P. Villamizar C., and René Gutiérrez Pérez. "X-ray structure analysis of symmetrically substituted 1,1′-diformylruthenocene." Acta Crystallographica Section E Crystallographic Communications 74, no. 9 (August 10, 2018): 1186–89. http://dx.doi.org/10.1107/s2056989018010642.

Full text
Abstract:
1,1′-Diformylruthenocene, [Ru(C6H5O)2], crystallizes in the orthorhombic system in the P212121 space group at room temperature. There are two crystallographically independent molecules in the asymmetric unit. The cyclopentadienyl rings have eclipsed configuration. The molecules self-assemble in a two-dimensional structure by C—H...O and C—H...π interactions with cisoid relative orientations of the two formyl groups. The crystal studied was refined as an inversion twin.
APA, Harvard, Vancouver, ISO, and other styles
3

Jagtap, Rohidas M., Shridhar H. Thorat, Rajesh G. Gonnade, Ayesha A. Khan, and Satish K. Pardeshi. "X-ray crystal structures and anti-breast cancer property of 3-tert-butoxycarbonyl-2-arylthiazolidine-4-carboxylic acids." New Journal of Chemistry 42, no. 2 (2018): 1078–86. http://dx.doi.org/10.1039/c7nj02961f.

Full text
Abstract:
The present article encompasses resolution and X-ray crystallographically confirmed absolute stereochemistry-correlated anticancer activity of diastereomeric 3-(tert-butoxycarbonyl)-2-(2-aryl)thiazolidine-4-carboxylic acids against MCF7 breast cancer cells.
APA, Harvard, Vancouver, ISO, and other styles
4

Smarra, A. L. S., R. K. Arni, W. F. De Azevedo, M. F. Colombo, and G. O. Bonilla-Rodriguez. "Crystallization and X-ray analysis of oxyhemoglobin-i liposarcus anisitsi (pisces)." Protein & Peptide Letters 4, no. 5 (October 1997): 349–54. http://dx.doi.org/10.2174/092986650405221017155228.

Full text
Abstract:
Abstract: Hemoglobin remains, despite the enormous amount of research involving this molecule, as a prototype for allosteric models and new conformations. Functional studies carried out on Hemoglobin-I from the South­ American Catfish Liposarcus anisitsi [1] suggest the existence of conformational states beyond those already described for human hemoglobin, which could be confirmed crystallographically. The present work represents the initial steps towards that goal.
APA, Harvard, Vancouver, ISO, and other styles
5

Chandran, Anu V., J. Rajan Prabu, G. P. Manjunath, K. Neelakanteshwar Patil, K. Muniyappa, and M. Vijayan. "Crystallization and preliminary X-ray studies of the C-terminal domain ofMycobacterium tuberculosisLexA." Acta Crystallographica Section F Structural Biology and Crystallization Communications 66, no. 9 (August 28, 2010): 1093–95. http://dx.doi.org/10.1107/s174430911003068x.

Full text
Abstract:
The C-terminal domain ofMycobacterium tuberculosisLexA has been crystallized in two different forms. The form 1 and form 2 crystals belonged to space groupsP3121 andP31, respectively. Form 1 contains one domain in the asymmetric unit, while form 2 contains six crystallographically independent domains. The structures have been solved by molecular replacement.
APA, Harvard, Vancouver, ISO, and other styles
6

Aitken, R., Neil Keddie, and Alexandra Slawin. "3,5-Dithiatricyclo[5.2.1.02,6]decan-4-one." Molbank 2020, no. 2 (April 24, 2020): M1126. http://dx.doi.org/10.3390/m1126.

Full text
Abstract:
The X-ray structure of the title compound has been determined and the structure shows an exo-configured planar dithiolanone ring. This is in contrast to the few previous dithiolanones to be characterised crystallographically, which are all twisted.
APA, Harvard, Vancouver, ISO, and other styles
7

Ozerov, Ruslan P., Victor A. Streltsov, Alexander N. Sobolev, Brian N. Figgis, and Victor L. Volkov. "Electron density in the sodium vanadium oxide bronze β-Na x V2O5 at 9 K." Acta Crystallographica Section B Structural Science 57, no. 3 (May 25, 2001): 244–50. http://dx.doi.org/10.1107/s0108768101002038.

Full text
Abstract:
The crystal structure and electron density in sodium vanadium oxide bronze, β-Na x V2O5 [x = 0.282 (3)], have been studied by accurate Mo Kα X-ray diffraction measurements at 9.6 (3) K. No noticeable difference in the crystal structures at room temperature and 9.6 K has been observed. No superstructure reflections, previously found by Kanai, Kagoshima & Nagasawa [(1982), J. Phys. Soc. Jpn, 51, 697–698], have been detected at low temperature. Analysis of the deformation electron density has revealed the presence of the quasi-two-dimensional sheets of the —V—O—V—O— bonds in the structure. The electron density in the different chemical bonds within each of the three crystallographically independent VO6 polyhedra noticeably varies, although there is no clear evidence that the three crystallographically independent V atoms have different valence states.
APA, Harvard, Vancouver, ISO, and other styles
8

Aitken, R. Alan, Charles Bloomfield, Liam J. R. McGeachie, and Alexandra M. Z. Slawin. "Diethyl pyrrole-2,5-dicarboxylate." Molbank 2020, no. 1 (February 17, 2020): M1117. http://dx.doi.org/10.3390/m1117.

Full text
Abstract:
The title compound was obtained in moderate yield by a new and unexpected base-induced ring contraction from a 1,4-thiazine precursor. Its X-ray structure showing hydrogen bonded dimers was compared with those of other crystallographically characterised 2-acylpyrroles.
APA, Harvard, Vancouver, ISO, and other styles
9

Weissbach, Torsten, Tilmann Leisegang, Andreas Kreyssig, Matthias Frontzek, Jens-Uwe Hoffmann, Dmitri Souptel, Anke Köhler, Günter Behr, Peter Paufler, and Dirk C. Meyer. "Intergrowth of several solid phases from the Y–Ni–B–C system in a large YNi2B2C crystal." Journal of Applied Crystallography 41, no. 4 (June 14, 2008): 738–46. http://dx.doi.org/10.1107/s002188980801279x.

Full text
Abstract:
A YNi2B2C single crystal containing traces of foreign phases was inspected by means of neutron and X-ray diffraction as well as scanning electron microscopy and X-ray spectroscopy methods. The diffraction patterns obtained from the experiments look similar to those expected for a superstructure. Nevertheless, they can be interpreted as crystallographically oriented precipitations of YB2C2and Ni2B within the YNi2B2C crystal, formed during the cooling process. The orientation relation between the lattices was obtained from experimental neutron and X-ray data. Structure refinements of the collected X-ray data were performed by separation of the intensity data of the individual phases. Scanning electron microscopy images of the inclusions found on a polished cross section of the crystal are presented; their chemical composition was determined using wavelength-dispersive X-ray analysis.
APA, Harvard, Vancouver, ISO, and other styles
10

Kim, Hyunjeong, Kouji Sakaki, Kohta Asano, Miho Yamauchi, Akihiko Machida, Tetsu Watanuki, and Yumiko Nakamura. "In-situ hydrogen gas loading setup for synchrotron X-ray total scattering." Acta Crystallographica Section A Foundations and Advances 70, a1 (August 5, 2014): C868. http://dx.doi.org/10.1107/s2053273314091311.

Full text
Abstract:
Hydrogen has been considered as a promising alternative fuel for transportation, provided we can find a way to store a large amount of hydrogen in a compact way. The realization of such a storage system can be achieved by developing materials that can easily absorb, safely store, and rapidly release hydrogen repeatedly. However, there is currently no material to meet all the requirements for on board storage. Great efforts have been made to understand hydrogenation properties of currently available materials to look for a way to improve properties or to prepare new materials. However, investigating the structure of some of these materials is challenging since their hydrides are only available under hydrogen gas pressure. Furthermore, many novel materials with improved properties often show heavily disordered or nanoscale structural features which are difficult to characterize using conventional crystallographic technique alone (crystallographically challenged hydrogen storage materials). In order to investigate the structural change in crystallographically challenged hydrogen storage materials during hydrogenation or dehydrogenation processes we have developed in-situ hydrogen gas loading setup for synchrotron X-ray total scattering experiments at the Japan Atomic Energy Agency (JAEA) beamline of BL22XU [1] at SPring-8. Coupled to an area detector [1,2], this setup allows us to obtain the atomic pair distribution function (PDF) [3] of metal hydrides either in equilibrium or in non-equilibrium state with hydrogen. In this poster, we will introduce our in-situ setup and present some preliminary results on AB5-type intermetallic compounds and Pd nanoparticles.
APA, Harvard, Vancouver, ISO, and other styles
11

Stachowski, Timothy R., Mary E. Snell, and Edward H. Snell. "SAXS studies of X-ray induced disulfide bond damage: Engineering high-resolution insight from a low-resolution technique." PLOS ONE 15, no. 11 (November 17, 2020): e0239702. http://dx.doi.org/10.1371/journal.pone.0239702.

Full text
Abstract:
A significant problem in biological X-ray crystallography is the radiation chemistry caused by the incident X-ray beam. This produces both global and site-specific damage. Site specific damage can misdirect the biological interpretation of the structural models produced. Cryo-cooling crystals has been successful in mitigating damage but not eliminating it altogether; however, cryo-cooling can be difficult in some cases and has also been shown to limit functionally relevant protein conformations. The doses used for X-ray crystallography are typically in the kilo-gray to mega-gray range. While disulfide bonds are among the most significantly affected species in proteins in the crystalline state at both cryogenic and higher temperatures, there is limited information on their response to low X-ray doses in solution, the details of which might inform biomedical applications of X-rays. In this work we engineered a protein that dimerizes through a susceptible disulfide bond to relate the radiation damage processes seen in cryo-cooled crystals to those closer to physiologic conditions. This approach enables a low-resolution technique, small angle X-ray scattering (SAXS), to detect and monitor a residue specific process. A dose dependent fragmentation of the engineered protein was seen that can be explained by a dimer to monomer transition through disulfide bond cleavage. This supports the crystallographically derived mechanism and demonstrates that results obtained crystallographically can be usefully extrapolated to physiologic conditions. Fragmentation was influenced by pH and the conformation of the dimer, providing information on mechanism and pointing to future routes for investigation and potential mitigation. The novel engineered protein approach to generate a large-scale change through a site-specific interaction represents a promising tool for advancing radiation damage studies under solution conditions.
APA, Harvard, Vancouver, ISO, and other styles
12

Kour, Dalbir, D. R. Patil, M. B. Deshmukh, Vivek K. Gupta, and Rajni Kant. "Synthesis and X-Ray Crystal Structure of Two Acridinedione Derivatives." Journal of Crystallography 2014 (February 27, 2014): 1–8. http://dx.doi.org/10.1155/2014/914504.

Full text
Abstract:
The two acridinedione derivatives 1 [3,3,6,6-tetramethyl-9-(4-methoxyphenyl)-3,4,6,7,9,10-hexahydro-2H,5H-acridine-1,8-dione (C24H29NO3)] and 2 [3,3,6,6-tetramethyl-9-(4-methylphenyl)-3,4,6,7,9,10-hexa-hydro-2H,5H-acridine-1,8-dione (C24H29NO2)] were synthesized and their crystal structures were determined by direct methods. The asymmetric unit of compound 1 contains two independent molecules. The 1,4-dihydropyridine (DHP) ring adopts boat conformation in both 1 and 2. In 1 the dione rings exist in sofa conformation (for both the crystallographically independent molecules) while the corresponding rings in 2 adopt half chair and sofa conformations, respectively. The crystal packing is stabilized by intermolecular N–H⋯O and C–H⋯O interactions in compound 1 and N–H⋯O interactions in compound 2.
APA, Harvard, Vancouver, ISO, and other styles
13

Corbey, Jordan F., Dallas D. Reilly, Lucas E. Sweet, and Timothy G. Lach. "Extraction of plutonium-containing microcrystals from Hanford soil using a focused ion beam for single-crystal X-ray diffraction analysis." Journal of Applied Crystallography 52, no. 6 (October 11, 2019): 1244–52. http://dx.doi.org/10.1107/s1600576719012299.

Full text
Abstract:
Herein, the successful use of a focused ion beam/scanning electron microscope to prepare microsamples of radioactive single crystals for X-ray diffraction analysis is reported. This technique was used to extract and analyze crystalline Pu-containing particles as small as 28 µm3 from Hanford soil taken from the 216-Z-9 waste crib, which were then crystallographically characterized using single-crystal X-ray diffraction to confirm the cubic structure of PuO2. As a systematic proof of concept, the technique was first tested using UO2 crystals milled into cubic shapes with approximate volumes of 4620, 1331, 125, 8 and 1 µm3, in order to empirically determine the crystal size limits for characterization by a laboratory-based diffractometer with a sealed tube Mo or Ag anode X-ray source and a charge-coupled device detector.
APA, Harvard, Vancouver, ISO, and other styles
14

Kuś, Piotr, Peter G. Jones, and Rafał Celiński. "Structure Elucidation of 2-Amino-5-phenyl-2-oxazolin-4-one (Pemoline) and X-Ray Structure of its Hydrolysis Product 5-Phenyl-oxazolidine-2,4- dione." Zeitschrift für Naturforschung B 60, no. 8 (August 1, 2005): 853–57. http://dx.doi.org/10.1515/znb-2005-0806.

Full text
Abstract:
In this study we compare spectroscopic properties of pemoline (2-amino-5-phenyl-2-oxazolin- 4-one) and its acid hydrolysis product 5-phenyl-oxazolidine-2,4-dione. Crystallization of pemoline from aqueous acetic acid gave single crystals of compound 2, the structure of which was determined by X-ray studies. All four crystallographically independent molecules form dimers linked by N-H···O = C hydrogen bonds.
APA, Harvard, Vancouver, ISO, and other styles
15

Richter, Carsten, Dmitri Novikov, Enver Mukhamedzhanov, Michail Borisov, Elena Ovchinnikova, Alexey Oreshko, Ksenia Akimova, et al. "Defect induced forbidden X-ray reflections in RbH2PO4." Acta Crystallographica Section A Foundations and Advances 70, a1 (August 5, 2014): C1529. http://dx.doi.org/10.1107/s2053273314084708.

Full text
Abstract:
Resonant X-ray diffraction was used to study the proton jumps in hydrogen-bonded rubidium dihydrogen phosphate (RDP) crystals. In the paraelectric RDP phase, hydrogen is delocalized between two crystallographically equivalent positions. At lower temperatures, this symmetry can be broken, which defines the processes that lead to the para- to ferroelectric phase transition. We have measured the energy spectra of the forbidden reflections 006 and 550 at incident radiation energies close to the Rb K-edge in a wide temperature range, down to the temperature of the ferroelectric phase transition. In the paraelectric phase we observed a growth of integrated intensity for both forbidden reflections with temperature. This behavior is opposite to conventional non-resonant Bragg reflections, where the intensity decreases in accordance with the Debye-Waller factor. The developed theoretical model explains this effect with the thermal motion induced (TMI) scattering mechanism and also confirms the adiabatic approximation stating that electrons instantly follow the nuclei movements. In the 550 energy spectra, we have observed an additional contribution to the resonant structure factor which could be associated with the presence of transient Slater-type proton configurations (PC) in the half-filled hydrogen position.
APA, Harvard, Vancouver, ISO, and other styles
16

Sharma, Varun, Indrajit Karmakar, Goutam Brahmachari, and Vivek Kumar Gupta. "X-ray crystal structure analysis of N'-acetyl-N'-phenyl-2-naphthohydrazide." European Journal of Chemistry 13, no. 3 (September 30, 2022): 253–58. http://dx.doi.org/10.5155/eurjchem.13.3.253-258.2235.

Full text
Abstract:
N'-Acetyl-N'-phenyl-2-naphthohydrazide, a biologically relevant organic molecule, was synthesized following a reported method and characterized based on its single X-ray crystallographic studies. The present manuscript deals with its detailed molecular interactions and X-ray crystal structure. Its space group is P-1 with the following unit cell parameters: a = 8.9164(7), b = 9.7058(9), c = 17.7384(12) Å, α = 88.308(7)°, β = 89.744(6)°, γ = 86.744(7)° and Z = 2. Crystal structure was solved by direct method and refined by full matrix least squares procedure to a final R value of 0.0580 and to a GOOF value of 1.066. The X-ray diffraction analyses showed that the asymmetric unit contains two crystallographically independent molecules. The crystal structure is stabilized by elaborate network of N-H···O and C-H···O hydrogen bonds along with C-H···π and π···π interactions to form supramolecular structures.
APA, Harvard, Vancouver, ISO, and other styles
17

Toyama, Mari, Tomoki Hasegawa, and Noriharu Nagao. "Colorimetric fluoride detection in dimethyl sulfoxide using a heteroleptic ruthenium(ii) complex with amino and amide groups: X-ray crystallographic and spectroscopic analyses." RSC Advances 12, no. 39 (2022): 25227–39. http://dx.doi.org/10.1039/d2ra03593f.

Full text
Abstract:
A bis-heteroleptic ruthenium(ii) complex; [Ru(Hdpa)2(H2pia)]X2 (1·X2; X = Cl, OTf, or F; Hdpa = di-2-pyridylamine; H2pia = 2-pycolinamide; OTf– = CF3SO3–) was synthesized and spectroscopically and crystallographically characterized.
APA, Harvard, Vancouver, ISO, and other styles
18

Hansen, Anna-Lena, Bastian Dietl, Martin Etter, Reinhard K. Kremer, David C. Johnson, and Wolfgang Bensch. "Temperature-dependent synchrotron X-ray diffraction, pair distribution function and susceptibility study on the layered compound CrTe3." Zeitschrift für Kristallographie - Crystalline Materials 233, no. 6 (June 27, 2018): 361–70. http://dx.doi.org/10.1515/zkri-2017-2100.

Full text
Abstract:
Abstract Results of combined synchrotron X-ray diffraction and pair distribution function experiments performed on the layered compound CrTe3 provide evidence for a short range structural distortion of one of the two crystallographically independent CrTe6 octahedra. The distortion is caused by higher mobility of one crystallographically distinct Te ion, leading to an unusual large Debye Waller factor. In situ high temperature X-ray diffraction investigations show an initial crystallization of a minor amount of elemental Te followed by decomposition of CrTe3 into Cr5Te8 and Te. Additional experiments provide evidence that the Te impurity (<1%) cannot be avoided. Analyses of structural changes in the temperature range 100–754 K show a pronounced anisotropic expansion of the lattice parameters. The differing behavior of the crystal axes is explained on the basis of structural distortions of the Cr4Te16 structural building units. An abrupt distortion of the structure occurs at T≈250 K, which then remains nearly constant down to 100 K. The structural distortion affects the spin exchange interactions between Cr3+ cations. A significant splitting between field-cooled (fc) and zero-field-cooled (zfc) magnetic susceptibility is observed below about 200 K. Applying a small external magnetic field results in a substantial spontaneous magnetization, reminiscent of ferro- or ferrimagnet exchange interactions below ~240 K. A Debye temperature of ~150 K was extracted from heat capacity measurements.
APA, Harvard, Vancouver, ISO, and other styles
19

Aramini, J., R. J. Batchelor, C. H. W. Jones, F. W. B. Einstein, and R. D. Sharma. "The X-ray crystal structure of bis(pentafluorophenyl)tellurium difluoride." Canadian Journal of Chemistry 65, no. 11 (November 1, 1987): 2643–48. http://dx.doi.org/10.1139/v87-437.

Full text
Abstract:
Bis(pentafluorophenyl)tellurium difluoride crystallizes, in the space group Cc, as a molecular solid in which pairs of crystallographically independent molecules are linked by two weak secondary [Formula: see text] interactions. (C12F12Te; a = 11.088(6), b = 20.040(10), c = 13.208(5); β = 109.07(4); U = 2774; Z = 8; fw = 499.7). The structural model was refined by full-matrix least-squares methods to a final residual of R1 = 0.035 for 1621 observed reflections. The stereochemistry of the primary bonding about Te is ψ-trigonal bipyramidal as for other R2TeX2 species and is compared with that of diphenyltellurium difluoride. Differences in the stereochemistry of the primary bonding about tellurium can be attributed to the influence of the more electron-withdrawing C6F5 group, and relates to the different secondary bonding and the crystal packing. 125Te nmr coupling constants, [Formula: see text], are reported for this and two related compounds. Previous 125Te Mössbauer data is discussed in the light of the structure.
APA, Harvard, Vancouver, ISO, and other styles
20

Banwell, Martin G., Bernard L. Flynn, Ernest Hamel, and Anthony C. Willis. "Synthesis, X-Ray Crystal Structure and Tubulin- Binding Properties of a Benzofuran Analogue of the Potent Cytotoxic Agent Combretastatin A4." Australian Journal of Chemistry 52, no. 8 (1999): 767. http://dx.doi.org/10.1071/ch99022.

Full text
Abstract:
The benzofuran (4), a ring-fused analogue of the potent antimitotic agent combretastatin A4 (1), has been prepared by a convergent route involving 5-endo-dig iodocyclization of o-hydroxytolan (5) as the key step. Compound (4), which has been characterized crystallographically as well as spectroscopically, is inactive as a tubulin-binding agent.
APA, Harvard, Vancouver, ISO, and other styles
21

Oyarzun, P., V. Gautier, M. Reich, and T. Vargas. "Rietveld Refinement of X-Ray Diffractograms Evidences Surface Texturization in Chemical and Biological Leaching of Chalcopyrite at 70°C." Advanced Materials Research 71-73 (May 2009): 389–92. http://dx.doi.org/10.4028/www.scientific.net/amr.71-73.389.

Full text
Abstract:
In the present work the occurrence of surface transformations triggered during chemical and biological leaching of chalcopyrite at 70 °C in basal medium were investigated with the aid of the Rietveld technique. Leaching experiments were conducted in 250 ml shake flasks at 150 rpm, contacting 1 g of a -80 # + 120# chalcopyrite concentrate with 100 ml of iron-free basal medium at pH 1.5 at 70°C. Three different conditions were used: experiment (a) in aerated conditions with basal medium inoculated with Sulfolobus metallicus (4x108 cells/ml); experiment (b) in aerated abiotic conditions; experiment (c) in abiotic conditions under N2 atmosphere. Copper and iron dissolved in solution was analyzed with atomic absorption. Samples of the initial chalcopyrite and the treated samples obtained after 16 days under the various experimental conditions were analyzed by Grazing-Incident X-ray diffraction (GID). The obtained x-ray spectra were then analyzed with the Rietveld refinement technique using the Topas software. Results showed that the initial concentrate consisted mainly of tetragonal chalcopyrite (JCPDS : 37-0471). Chalcopyrite leached under abiotic-anaerobic conditions was crystallographically unaffected. In chalcopyrite leached under aerated abiotic conditions there was evidence of surface texturization showed as preferential orientation to the crystallographical planes ( 3 1 2 ) and (1 1 6). Finally, surface texturization was also observed in chalcopyrite leached under inoculated conditions, but in this case showed as preferential orientation to the crystallographical planes ( 0 2 4 ) and ( 2 2 0). These results give a first evidence of surface texturization phenomena during dissolution of chalcopyrite at 70 °C. They suggest that chalcopyrite dissolution occurs via selective leaching of specific crystallographic planes, and this selectivity is influenced by both the chemical action of oxygen and the microorganism actvity.
APA, Harvard, Vancouver, ISO, and other styles
22

Brodersen, Klaus, and Karl Böhm. "Zur Koordination von Thiocyanat in ternären Metallkomplexen Die Kristallstruktur von Ba2Cd(SCN)6 • 7 H2O / On the Coordination of Thiocyanate in Ternary Metal Complexes The Crystal Structure of Ba2Cd(SCN)6 • 7 H2O." Zeitschrift für Naturforschung B 41, no. 4 (April 1, 1986): 439–43. http://dx.doi.org/10.1515/znb-1986-0408.

Full text
Abstract:
The title compound was prepared by reaction of aqueous solutions of Ba(SCN)2 • 3 H2O and CdSO4 • 8/3 H2O in the molar ratio of 3:1. The crystal structure was solved by X-ray methods (Mr = 861.61, space group P 21/c, Z = 4, a = 943.6(3) pm, b = 1469.7(6) pm, c = 1827.1(6) pm, β = 109.02(7), V = 2395.4 106 pm3, λ (AgKa) = 55.83 pm, dc = 2.38 gcm-3, μ(AgKa) = 22.4 cm-1, F(000) = 1591.9, T = 298 K, final R = 0.035 for 1350 independent reflexions). The structure consists of two crystallographically independent, distorted Cd(SCN)6 octahedra. Both crystallographically different Ba2+ cations are surrounded by 4 oxygen atoms of water and 5 nitrogen atoms of SCN, forming chains in the direction o f the b-axis
APA, Harvard, Vancouver, ISO, and other styles
23

Mockenhaupt, Christoph, Ralf Eßmann, and Heinz Dieter Lutz. "[Ni(NH3 )6]S04 : Kristallstruktur und Infrarotspektren / [Ni(NH3)6]SO4: Crystal Structure and Infrared Spectra." Zeitschrift für Naturforschung B 54, no. 7 (July 1, 1999): 843–48. http://dx.doi.org/10.1515/znb-1999-0704.

Full text
Abstract:
The crystal structure of [Ni(NH3)6]SO4 has been determined by single-crystal X-ray diffraction (P21/c; Z = 4; a = 705.0(1), b = 1195.2(2), c = 1180.0(2) pm, β = 96.14(3)°; 2271 reflections; R1 = 3.94%). In the hitherto unknown structure type, both the [Ni(NH3)6]2+ and the SO42- ions form chains along [100], The six crystallographically different ammine ligands of the distorted [Ni(NH3)6]2+ octahedra are involved in hydrogen bonds to six crystallographically equivalent SO42- ions (site symmetry C1). The strength of the hydrogen bonds differs strongly (νOD of matrix isolated NH2D molecules: 2378 - 2494 cm-1 , N ∙∙∙ O distances: 272 - 340 pm). The temperature evolution of the IR bands reveals the decrease of the dynamic orientational disorder of the NH3 molecules with decreasing temperature.
APA, Harvard, Vancouver, ISO, and other styles
24

Aatiq, Abderrahim, Btissame Haggouch, Rachid Bakri, Youssef Lakhdar, and Ismael Saadoune. "Structural characterization of two K2SnX(PO4)3 (X=Fe,Yb) with langbeinite structure." Powder Diffraction 21, no. 3 (September 2006): 214–19. http://dx.doi.org/10.1154/1.2246229.

Full text
Abstract:
Structures of two K2SnX(PO4)3(X=Fe,Yb) phosphates, obtained by conventional solid state reaction techniques at 950 °C, were determined at room temperature by X-ray powder diffraction using Rietveld analysis. The two materials exhibit the langbeinite-type structure (P213 space group, Z=4). Cubic unit cell parameter values are: a=9.9217(4) Å and a=10.1583(4) Å for K2SnFe(PO4)3 and K2SnYb(PO4)3, respectively. Structural refinements show that the two crystallographically independent octahedral sites (of symmetry 3) have a mixed Sn∕X (X=Fe,Yb) population although ordering is stronger in the Yb phase than in the Fe phase.
APA, Harvard, Vancouver, ISO, and other styles
25

Xu, Xiao-Juan. "A two-dimensional coordination polymer: poly[diaqua(μ4-4,4′-oxydibenzoato)(μ2-4,4′-oxydibenzoato)dilead(II)]." Acta Crystallographica Section C Structural Chemistry 70, no. 12 (November 6, 2014): 1105–8. http://dx.doi.org/10.1107/s2053229614023754.

Full text
Abstract:
In a new two-dimensional coordination polymer, [Pb(C14H8O5)(H2O)]n, the asymmetric unit consists of a PbIIcation, two halves of two crystallographically distinct fully deprotonated 4,4′-oxydibenzoate ligands and one aqua ligand. Single-crystal X-ray diffraction analysis reveals that the compound is a coordination polymer with the point symbol {53}2{54.82}. In addition, it exhibits a strong fluorescence emission in the solid state at room temperature.
APA, Harvard, Vancouver, ISO, and other styles
26

Hayton, Trevor W., Brian O. Patrick, Peter Legzdins, and W. Stephen McNeil. "The solid-state molecular structure of W(NO)3Cl3 and the nature of its W—NO bonding." Canadian Journal of Chemistry 82, no. 2 (February 1, 2004): 285–92. http://dx.doi.org/10.1139/v03-206.

Full text
Abstract:
The monomeric trinitrosyl complex, W(NO)3Cl3, can be prepared by the treatment of WCl6 in CH2Cl2 with NO gas, and its identity has been unambiguously confirmed by a single-crystal X-ray diffraction analysis. The complex crystallizes in the space group Pmn21 as a three-component twin (a = 10.4280(4) Å, b = 6.3289(2) Å, c = 5.6854(2) Å, Z = 2, R1 = 0.065, wR2 = 0.176). Its solid-state molecular structure consists of a tungsten centre bound to three chloride ligands and three linear nitrosyl ligands in a fac-octahedral stereochemistry. In addition, the structure contains a crystallographically imposed mirror plane. The two independent W—N linkages are 1.88(2) and 1.92(1) Å long, while the two corresponding N—O bond lengths are 1.13(2) and 1.16(2) Å. DFT calculations on fac-W(NO)3Cl3 at the B3LYP/LANL2DZ level of theory afford optimized intramolecular metrical parameters that match the X-ray crystallographically determined bond lengths and bond angles quite well. In addition, they provide a rationale for the nearly linear W-N-O linkages extant in the complex. Solutions of fac-W(NO)3Cl3 in CH2Cl2 lose ClNO under ambient conditions and deposit the well-known [W(NO)2Cl2]n polymer, and this conversion is fully reversible.Key words: nitrosyl, tungsten, structure, bonding.
APA, Harvard, Vancouver, ISO, and other styles
27

Corbett, M. H., G. Catalan, R. M. Bowman, and J. M. Gregg. "The effect of target crystallography on the growth of Pb(Mg1/3Nb2/3)O3 thin films using pulsed laser deposition." Journal of Materials Research 14, no. 6 (June 1999): 2355–58. http://dx.doi.org/10.1557/jmr.1999.0313.

Full text
Abstract:
Pulsed laser deposition has been used to make two sets of lead magnesium niobate thin films grown on single-crystal h100j MgO substrates. One set was fabricated using a perovskite-rich target while the other used a pyrochlore-rich target. It was found that the growth conditions required to produce almost 100% perovskite Pb(Mg1/3Nb2/3)O3 (PMN) films were largely independent of target crystallography. Films were characterized crystallographically using x-ray diffraction and plan view transmission electron microscopy, chemically using energy dispersive x-ray analysis, and electrically by fabricating a planar thin film capacitor structure and monitoring capacitance as a function of temperature. All characterization techniques indicated that perovskite PMN thin films had been successfully fabricated.
APA, Harvard, Vancouver, ISO, and other styles
28

Gao, Jian, Yu Yuan, Ai-Jun Cui, Feng Tian, Ming-Yang He, and Qun Chen. "A tetranuclear Sn(IV) 4-thiazolecarboxylate complex: synthesis, structure and catalytic behavior in the bulk ring-opening polymerization of glycolide." Zeitschrift für Naturforschung B 71, no. 8 (August 1, 2016): 843–48. http://dx.doi.org/10.1515/znb-2015-0211.

Full text
Abstract:
AbstractThe reaction of 4-thiazolecarboxylic acid (Htzc) and dimethyltin(IV) dichloride with NaOH in mixed MeOH-H2O solvent led to the formation of a new Sn(IV) complex (Me2Sn)4(μ3-O)2(tzc)4 (1). Its structure has been characterized by elemental analysis, IR spectroscopy and single crystal and powder X-ray diffraction. Single-crystal X-ray diffraction revealed that complex 1 crystallizes in the monoclinic P21/c space group with Z = 2 and has a tetranuclear structure with crystallographically imposed centrosymmetry. The as-synthesized complex 1 was found to be active toward the bulk solvent-free polymerization of glycolide, producing poly(glycolic acid) with a number-average molecular weight up to 55.5 kDa.
APA, Harvard, Vancouver, ISO, and other styles
29

Kenny, Peter W., Janet Newman, and Thomas S. Peat. "Nitrate in the active site of protein tyrosine phosphatase 1B is a putative mimetic of the transition state." Acta Crystallographica Section D Biological Crystallography 70, no. 2 (January 31, 2014): 565–71. http://dx.doi.org/10.1107/s1399004713031052.

Full text
Abstract:
The X-ray crystal structure of the complex of protein tyrosine phosphatase 1B with nitrate anion has been determined and modelled quantum-mechanically. Two protomers were present in the structure, one with the mechanistically important WPD loop closed and the other with this loop open. Nitrate was observed bound to each protomer, making close contacts with the S atom of the catalytic cysteine and a tyrosine residue from a crystallographically related protomer.
APA, Harvard, Vancouver, ISO, and other styles
30

Griese, Julia J., and Martin Högbom. "X-ray reduction correlates with soaking accessibility as judged from four non-crystallographically related diiron sites." Metallomics 4, no. 9 (2012): 894. http://dx.doi.org/10.1039/c2mt20080e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Toraya, H., N. Masciocchi, and W. Parrish. "Rietveld powder structure refinement of Na2Al2Ti6O16; Comparison of synchrotron radiation and conventional x-ray tube datasets." Journal of Materials Research 5, no. 7 (July 1990): 1538–43. http://dx.doi.org/10.1557/jmr.1990.1538.

Full text
Abstract:
The crystal structure of Na2Al2Ti6O16 was refined by the Rietveld method using synchrotron radiation and conventional x-ray powder data, and the agreement factors were Rp = 3.35%, Rwp = 4.30%, and RBragg = 6.39% for synchrotron data. The formula based on the chemical analysis and 16 O atoms is Na1.97Al1.82Ti6.15O16. The crystal data are monoclinic, C2/m, a = 12.1239(3) Å, b = 3.7749(1) Å, c = 6.4180(2) Å, β = 107.59(1)°, V = 280.00(4) Å3, Z = 1, and Dx = 3.82 g cm−3. The site occupancy refinement showed a partial ordering of Al3+ and Ti4+ ions in the two-crystallographically independent octahedral sites.
APA, Harvard, Vancouver, ISO, and other styles
32

Yahia, Hamdi Ben, Ute Ch Rodewald, Birgit Heying, Sarkarainadar Balamurugan, and Rainer Pöttgen. "The Solid Solution Lu2–xScxSiO5." Zeitschrift für Naturforschung B 66, no. 11 (November 1, 2011): 1183–87. http://dx.doi.org/10.1515/znb-2011-1115.

Full text
Abstract:
The monoclinic silicates Lu2SiO5 and Sc2SiO5 (Y2SiO5 type, space groupC2/c) forma solid solution Lu2−xScxSiO5. Samples with x = 0.5, 0.8, 1.0 were synthesized ceramically from Lu2O3, Sc2O3, and SiO2. The structures of three crystals with x = 0.88, 0.77, and 0.50 were refined on the basis of single-crystal X-ray diffraction data. The rare earth (RE) atoms occupy two crystallographically different 8ƒ sites with oxygen coordination numbers (CN) of 6 (RE2) and 7 (RE1). Refinements of the occupancy parameters showed Lu/Sc mixing for both sites with a strong preference of the smaller scandium atoms for CN6
APA, Harvard, Vancouver, ISO, and other styles
33

Shing, Ka-Pan, Qingyun Wan, Xiao-Yong Chang, and Chi-Ming Che. "The first crystallographically characterised ruthenium(vi) alkylimido porphyrin competent for aerobic epoxidation and hydrogen atom abstraction." Chemical Communications 56, no. 32 (2020): 4428–31. http://dx.doi.org/10.1039/c9cc09972g.

Full text
Abstract:
The syntheses of [RuVI(Por)(NAd)(O)] and [RuVI(2,6-F2-TPP)(NAd)2] have been described. [RuVI(Por)(NAd)(O)] characterised by X-ray crystallography is competent for H-atom abstraction and aerobic alkene epoxidation.
APA, Harvard, Vancouver, ISO, and other styles
34

Yan, Hong, Hui-Qin Wang, Cheng-Liang Ni, and Xiu-Qing Song. "A photodimer of a 4-phenyl-4H-pyran." Acta Crystallographica Section E Structure Reports Online 62, no. 5 (April 21, 2006): o1951—o1953. http://dx.doi.org/10.1107/s1600536806012980.

Full text
Abstract:
A new cage photodimer, tetraethyl 2,4,8,10-tetramethyl-6,12-diphenyl-3,9-dioxapentacyclo[6.4.0.02,7.04,11.05,10]dodecane-1,5,7,11-tetracarboxylate, C38H44O10, was prepared through [2+2]-photocycloaddition of diethyl 2,6-dimethyl-4-phenyl-4H-pyran-3,5-dicarboxylate in the solid state. The molecular structure was elucidated by X-ray diffraction analysis, 1H NMR, IR and mass spectroscopy, and elemental analysis. The molecule possesses a crystallographically imposed centre of symmetry. The crystal structure is stabilized by weak C—H...O hydrogen-bond interactions.
APA, Harvard, Vancouver, ISO, and other styles
35

Seidel, Rüdiger W., and Richard Goddard. "Anisole at 100 K: the first crystal structure determination." Acta Crystallographica Section C Structural Chemistry 71, no. 8 (July 8, 2015): 664–66. http://dx.doi.org/10.1107/s2053229615012553.

Full text
Abstract:
The simplest alkyl aryl ether, anisole (methoxybenzene), C7H8O, is a feedstock chemical and is widely used in the pharmaceutical industry. The structure of anisole at 100 K, as determined by single-crystal X-ray analysis, is reported. A crystal (m.p. 236 K) suitable for X-ray diffraction was obtained from the melt. The title compound crystallizes in the centrosymmetric space groupP21/cwith two molecules in the asymmetric unit (Z′ = 2). Both crystallographically distinct molecules adopt a virtually flat (Cs-symmetric) conformation. The arrangement of the molecules in the solid state appears to be governed by close packing. No face-to-face π–π stacking of the molecules is observed, but rather edge-to-face interactions result in a herringbone packing motif.
APA, Harvard, Vancouver, ISO, and other styles
36

Waychunas, G. A. "Natural nanoparticle structure, properties and reactivity from X-ray studies." Powder Diffraction 24, no. 2 (June 2009): 89–93. http://dx.doi.org/10.1154/1.3132590.

Full text
Abstract:
Synthetic analogs of naturally occurring nanoparticles have been studied by a range of X-ray techniques to determine their structure and chemistry, and relate these to their novel chemical properties and physical behavior. ZnS nanoparticles, formed in large concentrations naturally bymicrobial action, have an interesting core-shell structure with a highly distorted and strained outer layer. The strain propagates through the particles and produces unusual stiffness but can be relieved by changing the nature of the surface ligand binding. Weaker bound ligands allow high surface distortion, but strongly bound ligands relax this structure and reduce the overall strain. Only small amounts of ligand exchange causes transformations from the strained to the relaxed state. Most remarkably, minor point contacts between strained nanoparticles also relax the strain. Fe oxyhydroxide nanoparticles appear to go through structural transformations dependent on their size and formation conditions, and display a crystallographically oriented form of aggregation at the nanoscale that alters growth kinetics. At least one Fe oxyhydroxide mineral may only be stable on the nanoscale, and nonstoichiometry observed on the hematite surface suggests that for this phase and possibly other natural metal oxides, chemistry may be size dependent. Numerous questions exist on nanominerals formed in acid mine drainage sites and by reactions at interfaces.
APA, Harvard, Vancouver, ISO, and other styles
37

Romero, Sergio Antonio, Christien G. Hauegen, Fernando J. G. Landgraf, and Marcos Flavio de Campos. "EBSD Analysis of SmCoFeCuZr Alloys." Materials Science Forum 869 (August 2016): 608–13. http://dx.doi.org/10.4028/www.scientific.net/msf.869.608.

Full text
Abstract:
In the present study, EBSD was used for the characterization of alloys used for production of SmFeCoCuZr magnets. EBSD is adequate for texture analysis, but may give misleading results for phase identification. EBSD is not suitable for identifying phases with very similar crystalline structure, especially when the phases are crystallographically coherent, due to the superposition of Kikuchi lines. As consequence, for phase identification EBSD should be considered a complementary technique to other methods, as for example x-ray diffraction (XRD).
APA, Harvard, Vancouver, ISO, and other styles
38

Mattei, Carlo Andrea, Bertrand Lefeuvre, Vincent Dorcet, Gilles Argouarch, Olivier Cador, Claudia Lalli, and Fabrice Pointillart. "Counterintuitive Single-Molecule Magnet Behaviour in Two Polymorphs of One-Dimensional Compounds Involving Chiral BINOL-Derived Bisphosphate Ligands." Magnetochemistry 7, no. 11 (November 16, 2021): 150. http://dx.doi.org/10.3390/magnetochemistry7110150.

Full text
Abstract:
The coordination reaction of the [Dy(hfac)3(H2O)2] units (hfac− = 1,1,1,5,5,5-hexafluoroacetylacetonate) with the [8′-(Diphenoxylphosphinyl)[1,1′-binaphthalen]-8-yl]diphenoxylphosphine oxide ligand (L) followed by a crystallisation in a 1:3 CH2Cl2:n-hexane solvent mixture led to the isolation of a new polymorph of formula [(Dy(hfac)3((S)-L))3]n (1). The X-ray structure on single crystal of 1 revealed the formation of a mono-dimensional coordination polymer with three crystallographically independent DyIII centres, which crystallised in the polar chiral P21 space group. Ac magnetic measurements highlighted single-molecule magnet behaviour under both zero and 1000 Oe applied magnetic field with magnetic relaxation through quantum tunneling of the magnetisation (QTM, zero field only) and Raman processes. Despite the three crystallographically independent DyIII centres adopting a distorted D4d coordination environment, a single slow magnetic relaxation contribution was observed at a slower rate than its previously studied [(Dy(hfac)3((S)-L))]n (2) polymorph.
APA, Harvard, Vancouver, ISO, and other styles
39

Hamada, E., N. Ishizawa, F. Marumo, K. Ohsumi, Y. Shimizugawa, K. Reizen, and T. Matsunami. "Structure of Mg6SO2(OH)14 determined by micro single-crystal X-ray diffraction." Acta Crystallographica Section B Structural Science 52, no. 2 (April 1, 1996): 266–69. http://dx.doi.org/10.1107/s0108768195014443.

Full text
Abstract:
The structure of hexamagnesium sulfonyl tetradecahydroxide, Mg6SO2(OH)14, has been determined with an extremely small single-crystal (ca 0.5 × 100 × 2.5 μm) using synchrotron radiation [1.00 (1) Å]. The crystal is orthorhombic with space group Ccmm (No. 63), a = 15.895 (1), b = 3.105 (1), c = 13.367 (1) Å, V = 659.64 (3) Å3 and Z = 2. The final R and wR values are 0.073 and 0.083 for 105 crystallographically independent reflections. The structure consists of MO6 octahedra and SO4 tetrahedra. The MgO6 octahedra share their edges to form zigzag sheets parallel to (100). The SO4 tetrahedron shares two of the O atoms with the MgO6 octahedra at the turning of the sheet and is statistically located with a probability of 0.5. Fine streaks observed perpendicular to the b* axis on diffraction photographs indicate that the SO4 tetrahedra are distributed at random along the a and c axes, but alternately along the b axis. All the H atoms are considered to exist in the form of OH groups.
APA, Harvard, Vancouver, ISO, and other styles
40

Munro, Orde Quentin, and Lynette Mariah. "Conformational analysis: crystallographic, molecular mechanics and quantum chemical studies of C—H...O hydrogen bonding in the flexible bis(nosylate) derivative of catechol." Acta Crystallographica Section B Structural Science 60, no. 5 (September 15, 2004): 598–608. http://dx.doi.org/10.1107/s0108768104019846.

Full text
Abstract:
The single-crystal X-ray diffraction analysis of 2-{[(4-nitrophenoxy)sulfonyl]oxy}phenyl 4-nitrophenyl sulfate (4) reveals that an interesting intermolecular or extended structure (a one-dimensional hydrogen-bonded polymer) is formed because of pairs of intermolecular (aryl)C—H...O(nitro) hydrogen bonds between the C 2 symmetry monomer units. The axis of the hydrogen-bonded polymer runs co-linear with the [101] face diagonal of the monoclinic unit cell. Molecular mechanics calculations using a modified version of the MM+ force field and a random conformational search algorithm have been used to locate the important low-energy in vacuo conformations of (4). The MM-calculated conformation of (4) that most closely matches the X-ray structure lies some 26.5 kJ mol−1 higher in energy than the global minimum-energy conformation, consistent with the notion that the crystallographically observed molecular architecture of (4) is a local energy minimum in the absence of its crystal lattice environment. Since the X-ray conformation of (4) was correctly calculated only when all of the neighbouring molecules in the crystal lattice were included in the simulation, hydrogen bonding and other non-bonded interactions in the crystal lattice clearly dictate the experimentally observed conformation of (4). Quantum chemical calculations (AM1 method) confirm the critical role played by the intermolecular (aryl)C—H...O(nitro) hydrogen bonds in controlling the crystallographically observed conformation of (4) and show that self-recognition in this system by hydrogen bonding is favoured on electrostatic grounds. Collectively, the molecular simulations suggest that because the lowest-energy molecular conformation of (4) does not permit the formation of an extended hydrogen-bonded `supramolecular' structure, it is not the preferred conformation in the crystalline solid state.
APA, Harvard, Vancouver, ISO, and other styles
41

Waldhart, Greyson W., and Neal P. Mankad. "trans-Tetracarbonylbis(triphenylphosphane-κP)molybdenum(0)." Acta Crystallographica Section E Structure Reports Online 70, no. 2 (January 11, 2014): m36. http://dx.doi.org/10.1107/s1600536814000300.

Full text
Abstract:
The well known title compound,trans-[Mo(C18H15P)2(CO)4], has not been studied previously by X-ray crystallography, unlike itscisisomer. The complex possesses crystallographically imposed inversion symmetry, with the Mo atom residing on an inversion centre (1aWyckoff position). The two triphenylphosphane groups are arranged in a staggered orientation. Each of the phenyl groups exhibits significantly different Mo—P—C—C torsion angles ranging from 2.6 (2) to 179.4 (1)°, most likely due to steric interactions based upon their positions relative to the carbonyl ligands.
APA, Harvard, Vancouver, ISO, and other styles
42

Dickie, Diane A., Hanifa Jalali, Rahul G. Samant, Michael C. Jennings, and Jason AC Clyburne. "Synthesis and structural characterization of m-terphenyl Schiff base ligands and their aluminum complexes." Canadian Journal of Chemistry 82, no. 9 (September 1, 2004): 1346–52. http://dx.doi.org/10.1139/v04-110.

Full text
Abstract:
2,4,6-Triphenylbenzaldehyde 1 undergoes a condensation reaction with 2-aminophenol to give N-(2′,4′,6′-triphenylbenzylidene)-2-iminophenol (TPIP) 2. The imine 2 can be reduced with NaBH4 in ethanol to form N-(2′,4′,6′-triphenylbenzyl)-2-aminophenol (TPAP) 3. Addition of trimethylaluminum to 2 or 3 results in the formation of the complexes TPIP-AlMe2·AlMe3 (4) or TPAP-AlMe2 (5). Compounds 2, 3, and 4 have been crystallographically characterized.Key words: N,O ligands, aluminum, m-terphenyl, Schiff bases, X-ray crystallography.
APA, Harvard, Vancouver, ISO, and other styles
43

Demant, Udo, Elke Conradi, Ulrich Müller, and Kurt Dehnicke. "Formamidinhim-Hexachloroferrat(III) Synthese und Kristallstruktur / Formamidinium-Hexachloroferrate(III) Synthesis and Crystal Structure." Zeitschrift für Naturforschung B 40, no. 3 (March 1, 1985): 443–46. http://dx.doi.org/10.1515/znb-1985-0324.

Full text
Abstract:
[HC(NH2)2]3FeCl6 was obtained together with other products from the reaction of S4N4 with HCl in H2CCl2 in the presence of FeCl3. Its crystal structure was determined from X-ray diffraction data (473 independent observed reflexions, R = 0.047). Lattice constants: a = 961.6, c = 876.4 pm; tetragonal, space group P42/m, Z = 2. Of the two crystallographically independent formamidinium ions HC(NH2)2⊕, one exhibits positional disorder; the other one has C-N bond lengths of 128 pm. The FeCl63⊖ ions have symmetry C2h, but the deviation from Oh is small.
APA, Harvard, Vancouver, ISO, and other styles
44

Deflon, Victor M., Cassia C. de Sousa Lopes, Karl E. Bessler, Lincoln L. Romualdo, and Elke Niquet. "Preparation, Characterization and Crystal Structure of Lead(II) Tricyanomethanide." Zeitschrift für Naturforschung B 61, no. 1 (January 1, 2006): 33–36. http://dx.doi.org/10.1515/znb-2006-0107.

Full text
Abstract:
The so far unknown lead tricyanomethanide, Pb[C(CN)3]2, was obtained from a saturated aqueous solution of PbCl2 and solid AgC(CN)3. Its IR spectrum and thermal behaviour are described. The crystal structure was determined by single-crystal X-ray diffraction (trigonal, P31m, Z = 3, a = 1414.4(5), c = 409.02(6) pm, R1 = 0.0249, wR2 = 0.0527). Two crystallographically independent ninefold coordinated Pb atoms are connected by planar tricyanomethanide ions in two distinct bridging coordination modes. The Pb−N distances range between 254 and 293 pm.
APA, Harvard, Vancouver, ISO, and other styles
45

Steyl, Gideon, Leo Kirsten, Alfred Muller, and Andreas Roodt. "trans-Bromocarbonylbis(triphenylphosphine)rhodium." Acta Crystallographica Section E Structure Reports Online 62, no. 5 (April 26, 2006): m1127—m1129. http://dx.doi.org/10.1107/s1600536806014619.

Full text
Abstract:
The title compound, [RhBr(C18H15P)2(CO)], can be characterized as a rhodium(I) Vaska-type compound based only on the spectroscopic [νKBr(CO), 31P] data. A low-temperature X-ray crystallographic analysis shows that the compound possesses a crystallographically imposed centre of symmetry with a statistically disordered Br atom and CO group. The essentially different occupancies for the Br atom and the CO group [0.283 (2) and 0.717 (2), respectively] suggest that the compound exists as a mixture of rhodium(0), rhodium(I) and rhodium(II) complexes.
APA, Harvard, Vancouver, ISO, and other styles
46

Shima, Kaori, Naoki Mitsugi, and Hirotoshi Nagata. "Surface precipitates on single crystal LiNbO3 after dry-etching by CHF3 plasma." Journal of Materials Research 13, no. 3 (March 1998): 527–29. http://dx.doi.org/10.1557/jmr.1998.0068.

Full text
Abstract:
The CHF3 electron cyclotron resonance (ECR) plasma etched LiNbO3 (LN) surface was analyzed chemically and crystallographically to investigate the dry-etch machining process for LN crystal, which was recently needed to obtain broader-band optical modulators. The etched surface was entirely covered with amorphous-like precipitates having ~70 nm diameter. These precipitates (or a part of them) were thought to be LiF from Auger electron and x-ray photoelectron spectroscopy. The results indicated that the LiF was formed and remained on the etched surface while the Nb was almost completely removed.
APA, Harvard, Vancouver, ISO, and other styles
47

Pfannenschmidt, Ulrike, Ute Ch Rodewald, and Rainer Pöttgen. "ScIrP with ZrNiAl-type Structure." Zeitschrift für Naturforschung B 66, no. 2 (February 1, 2011): 205–8. http://dx.doi.org/10.1515/znb-2011-0214.

Full text
Abstract:
The phosphide ScIrP was synthesized from the elements in a bismuth flux and characterized by powder and singlecrystal X-ray diffraction: ZrNiAl type, P6̄2m, Z = 3, a = 637.2(3), c = 389.2(2) pm, wR2 = 0.0280, 250 F2 values, 15 variables. The two crystallographically independent phosphorus sites have tricapped trigonal-prismatic metal coordination P1Ir3Sc6 and P2Ir6Sc3. The shortest interatomic distances occur for Ir-P (244 - 251 pm) within the 3D [IrP] network in which the scandium atoms fill cavities of coordination number 15 (4 Sc + 6 Ir + 5 P).
APA, Harvard, Vancouver, ISO, and other styles
48

Hill, Sarah C., Daniel S. Jones, and Daniel Rabinovich. "[Bis(3-methyl-2-thioxo-2,3-dihydro-1H-imidazolyl)borato-κ2 S,S′]dibromoindium(III)." Acta Crystallographica Section E Structure Reports Online 62, no. 4 (March 8, 2006): m702—m704. http://dx.doi.org/10.1107/s1600536806007732.

Full text
Abstract:
The structure of the title compound, [InBr2(C8H12N4BS2)], the first bis(mercaptoimidazolyl)borate (BmMe) complex of indium to be structurally characterized, has been determined by single-crystal X-ray diffraction. The four-coordinate In atom displays a distorted tetrahedral geometry in the solid state and is surrounded by the two thione groups of a BmMe ligand and two bromides, with an average In—Br bond distance of 2.494 Å. The presence of a crystallographically imposed mirror plane results in the observation of a unique In—S bond length of 2.4407 (11) Å.
APA, Harvard, Vancouver, ISO, and other styles
49

Zhong, Kai-Long, Jing Quan, Xian-Xiao Pan, Wei Song, and Bing-Feng Li. "Synthesis, crystal structure and properties of a 2-D Cd(II) coordination polymer based on ferrocenecarboxylate and 4,4′-bipyridine ligands." Zeitschrift für Naturforschung B 77, no. 2-3 (December 16, 2021): 125–29. http://dx.doi.org/10.1515/znb-2021-0171.

Full text
Abstract:
Abstract A new cadmium(II)-based coordination polymer [Cd3(FcCOO)6(4,4′-bipy)(H2O)2] n (FcCOO = ferrocenecarboxylato and 4,4′-bipy = 4,4′-bipyridine) has been synthesized under hydrothermal conditions and characterized by single-crystal X-ray diffraction. The results of a crystal structural analysis has revealed that the title compound consists of two crystallographically unique CdII centers, one in a general position with a five-coordinated and one on an inversion center with a six-coordinated environment. The CdII centers are connected by FcCOO− units to form a metal carboxylate oxygen chain extending parallel to the [100] direction while the 4,4′-bipy ligands further act as bridging linkers of the CdII centers resulting in a layered polymer. In addition, an X-ray powder diffraction and thermal gravimetric analysis and a cyclo-voltammetric characterization of the complex have also been carried out.
APA, Harvard, Vancouver, ISO, and other styles
50

Lissner, Falk, Frank A. Weber, and Thomas Schleid. "Drei Formen des Samarium(III)-Sulfidselenids Sm2S2-xSe1+x (0,1 ≤ x < 0,2) / Tree Types of the Samarium(III) Sulfide Selenide Sm2S2-xSe1+x (0,1 ≤ x < 0,2)." Zeitschrift für Naturforschung B 56, no. 10 (October 1, 2001): 990–96. http://dx.doi.org/10.1515/znb-2001-1005.

Full text
Abstract:
Single crystals of Sm2S2-xSe1+x (0.1 ≤ x < 0.2) have been obtained for the first time through the oxidation of KSm2Cl5 with an excess of sulfur and selenium in equimolar amounts at 850 °C after seven days in evacuated silica tubes. They emerged as almost black, in thin layers deep red, lath-shaped needles (A and U type, respectively) as well as red, bead-shaped polyhedra (C type) of the gross chemical composition Sm2S2Se according to X-ray structure analyses. A much simpler method of synthesis is based on the direct fusion of the elements (samarium, sulfur and selenium) in appropriate molar ratios (2:2:1) in the presence of NaCl as a flux under otherwise analogous conditions (silica tubes, 7 d, 850 °C). A-Sm2S1.82Se1.18 crystallizes orthorhombically (a = 753.1(3), b = 401.9(1), c = 1565.8(6) pm, Z = 4) in the space group Pnma with the α-Gd2S3-type structure. Two crystallographically different Sm3+ cations are coordinated by eight (Sm1) and seven (Sm2) Ch2- anions (S2- and Se2-) as biand monocapped trigonal prisms, respectively. C-Sm2S1.90Se1.10 adopts the cubic γ-Ce2S3-type structure (I4̅3d; a = 858.7(2) pm, Z = 5.333) with trigon-dodecahedrally coordinated Sm3+ cations (CN = 8 ). Finally, U-Sm2S1.84Se1.16 with the orthorhombic U2S3-type structure (Pnma; a = 1105.3(6), b = 399.2(1), c = 1074.0(5) pm, Z = 4) exhibits two crystallographically different Sm3+ cations again, which are coordinated by seven (Sm1) and seven plus one (Sm2) Ch2- anions, respectively. The preferential occupation of S _ and Se2- anions at only one in the C-type, but three anionic sites each in the A- and U-type crystal structures of Sm2S2-xSe1+x (0 .1 ≤ x < 0.2 ) is discussed.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography