Journal articles on the topic 'Lanthanide Chemistry'

To see the other types of publications on this topic, follow the link: Lanthanide Chemistry.

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 'Lanthanide Chemistry.'

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

Evans, William J., and David S. Lee. "Early developments in lanthanide-based dinitrogen reduction chemistry." Canadian Journal of Chemistry 83, no. 4 (April 1, 2005): 375–84. http://dx.doi.org/10.1139/v05-014.

Full text
Abstract:
Although the first crystallographically characterized lanthanide dinitrogen complex was reported in 1988 with samarium, it is only in recent years that this field has expanded to include fully characterized examples for the entire series of lanthanides. The development of lanthanide dinitrogen chemistry has been aided by a series of unexpected results that present some good lessons in the development of science. This review presents a chronological account of the lanthanide dinitrogen chemistry discovered in our laboratory through the summer of 2004.Key words: lanthanides, dinitrogen, reduction, alkali metal, nitrogen fixation, diazenido.
APA, Harvard, Vancouver, ISO, and other styles
2

Camp, Clément, Valentin Guidal, Biplab Biswas, Jacques Pécaut, Lionel Dubois, and Marinella Mazzanti. "Multielectron redox chemistry of lanthanide Schiff-base complexes." Chemical Science 3, no. 8 (2012): 2433–48. http://dx.doi.org/10.1039/c2sc20476b.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Liu, Juewen. "Lanthanide-dependent RNA-cleaving DNAzymes as metal biosensors." Canadian Journal of Chemistry 93, no. 3 (March 2015): 273–78. http://dx.doi.org/10.1139/cjc-2014-0465.

Full text
Abstract:
Lanthanides represent a group of very important but challenging analytes for biosensor development. These 15 elements are very similar in their chemical properties. So far, limited success has been realized using the rational ligand design approach. My laboratory has successfully accomplished the task of carrying out combinatorial selection to isolate lanthanide-dependent RNA-cleaving DNAzymes. We report two new DNAzymes, each discovered in a different selection condition and both are highly specific to lanthanides. When both DNAzymes are used together, it is possible to identify the last few heavy lanthanides. Upon introducing a phosphorothioate modification, one of the abovementioned DNAzymes becomes highly active with many toxic heavy metals. With the selection of more DNAzymes with different activity patterns cross the lanthanide series, a sensor array might be produced for identifying each ion. This article is a minireview of the current developments on this topic and some of the historical aspects. It reflects the main content of the Fred Beamish Award presentation delivered at the 2014 Canadian Society for Chemistry Conference in Vancouver. Future directions in this area are also discussed.
APA, Harvard, Vancouver, ISO, and other styles
4

Weißhoff, Hardy, Katharina Janek, Peter Henklein, Herbert Schumann, and Clemens Mügge. "Elution Behavior and Structural Characterization of N- and C-functionalized DOTA Complexes for the Labelling of Biomolecules." Zeitschrift für Naturforschung B 64, no. 10 (October 1, 2009): 1159–68. http://dx.doi.org/10.1515/znb-2009-1008.

Full text
Abstract:
Two types of lanthanide complexes of 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid (DOTA) for the labelling of biomolecules were investigated by HPLC, MS and NMR spectroscopy. The elution behavior of lanthanide complexes of N-functionalized DOTA [1,4,7,10-tetraazacyclododecane- 1,4,7-triacetic acid-10-maleimidoethylacetamide (nDOTA-Mal) and 1-{2-[4-(maleimido- N-propylacetamidobutyl)amino]-2-oxoethyl}-1,4,7,10-tetraazacyclododecane-4,7,10-triacetic acid (nDOTA-Bu-Mal)] and C-functionalized DOTA [2-{4-(maleimido-N-propylacetamido)benzyl}-1,4, 7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid (cDOTA-Bnz-Mal) and 2-(4-isothiocyanatobenzyl)- 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid (cDOTA-Bnz-NCS)] was compared. N-functionalized lanthanide DOTA complexes coelute as required for their use as ICAT-analogous reagents. The complexation of the C-functionalized DOTA with lanthanides results in two fractions separable by HPLC. Coelution is observed for the main fractions of the lanthanide complexes. The retention times of the minor fractions show a dependence on the ionic radii of the metal ions. MALDI spectra of lanthanide-DOTA-peptide conjugates including different monoisotopic lanthanides demonstrate the advantage of the mass variations for extensive peptide and protein investigations.
APA, Harvard, Vancouver, ISO, and other styles
5

Savić, Aleksandar, Anna M. Kaczmarek, Rik Van Deun, and Kristof Van Hecke. "DNA Intercalating Near-Infrared Luminescent Lanthanide Complexes Containing Dipyrido[3,2-a:2′,3′-c]phenazine (dppz) Ligands: Synthesis, Crystal Structures, Stability, Luminescence Properties and CT-DNA Interaction." Molecules 25, no. 22 (November 13, 2020): 5309. http://dx.doi.org/10.3390/molecules25225309.

Full text
Abstract:
In order to create near-infrared (NIR) luminescent lanthanide complexes suitable for DNA-interaction, novel lanthanide dppz complexes with general formula [Ln(NO3)3(dppz)2] (Ln = Nd3+, Er3+ and Yb3+; dppz = dipyrido[3,2-a:2′,3′-c]phenazine) were synthesized, characterized and their luminescence properties were investigated. In addition, analogous compounds with other lanthanide ions (Ln = Ce3+, Pr3+, Sm3+, Eu3+, Tb3+, Dy3+, Ho3+, Tm3+, Lu3+) were prepared. All complexes were characterized by IR spectroscopy and elemental analysis. Single-crystal X-ray diffraction analysis of the complexes (Ln = La3+, Ce3+, Pr3+, Nd3+, Eu3+, Er3+, Yb3+, Lu3+) showed that the lanthanide’s first coordination sphere can be described as a bicapped dodecahedron, made up of two bidentate dppz ligands and three bidentate-coordinating nitrate anions. Efficient energy transfer was observed from the dppz ligand to the lanthanide ion (Nd3+, Er3+ and Yb3+), while relatively high luminescence lifetimes were detected for these complexes. In their excitation spectra, the maximum of the strong broad band is located at around 385 nm and this wavelength was further used for excitation of the chosen complexes. In their emission spectra, the following characteristic NIR emission peaks were observed: for a) Nd3+: 4F3/2 → 4I9/2 (870.8 nm), 4F3/2 → 4I11/2 (1052.7 nm) and 4F3/2 → 4I13/2 (1334.5 nm); b) Er3+: 4I13/2 → 4I15/2 (1529.0 nm) c) Yb3+: 2F5/2 → 2F7/2 (977.6 nm). While its low triplet energy level is ideally suited for efficient sensitization of Nd3+ and Er3+, the dppz ligand is considered not favorable as a sensitizer for most of the visible emitting lanthanide ions, due to its low-lying triplet level, which is too low for the accepting levels of most visible emitting lanthanides. Furthermore, the DNA intercalation ability of the [Nd(NO3)3(dppz)2] complex with calf thymus DNA (CT-DNA) was confirmed using fluorescence spectroscopy.
APA, Harvard, Vancouver, ISO, and other styles
6

Xu, Hengbin, Jiamiao Qu, Milin Zhang, Yongde Yan, Xin Sun, Yanghai Zheng, Min Qiu, and Li Liu. "The linear relationship derived from the deposition potential of Pb–Ln alloy and atomic radius." New Journal of Chemistry 42, no. 20 (2018): 16533–41. http://dx.doi.org/10.1039/c8nj03342k.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Pereira, Cláudia C. L., José M. Carretas, Bernardo Monteiro, and João P. Leal. "Luminescent Ln-Ionic Liquids beyond Europium." Molecules 26, no. 16 (August 10, 2021): 4834. http://dx.doi.org/10.3390/molecules26164834.

Full text
Abstract:
Searching in the Web of Knowledge for “ionic liquids” AND “luminescence” AND “lanthanide”, around 260 entries can be found, of which a considerable number refer solely or primarily to europium (90%, ~234). Europium has been deemed the best lanthanide for luminescent applications, mainly due to its efficiency in sensitization, longest decay times, and the ability to use its luminescence spectra to probe the coordination geometry around the metal. The remaining lanthanides can also be of crucial importance due to their different colors, sensitivity, and capability as probes. In this manuscript, we intend to shed some light on the existing published work on the remaining lanthanides. In some cases, they appear in papers with europium, but frequently in a subordinate position, and in fewer cases then the main protagonist of the study. All of them will be assessed and presented in a concise manner; they will be divided into two main categories: lanthanide compounds dissolved in ionic liquids, and lanthanide-based ionic liquids. Finally, some analysis of future trends is carried out highlighting some future promising fields, such as ionogels.
APA, Harvard, Vancouver, ISO, and other styles
8

Vassiliev, Valery P., Valery A. Lysenko, and Marcelle Gaune-Escard. "Relationship of thermodynamic data with Periodic Law." Pure and Applied Chemistry 91, no. 6 (June 26, 2019): 879–93. http://dx.doi.org/10.1515/pac-2018-0717.

Full text
Abstract:
Abstract Relationships between the various physical and chemical properties of isostructural compounds take place according to the Periodic Table that is a fundamental basis of Chemistry. The systematization of this approach, data vs. the Periodic Table, will contribute to further development of the solid state chemistry theory. The lanthanides and the actinides make up the f block of the Periodic Table. The lanthanides are the elements produced as the 4f sublevel is filled with electrons and the actinides are formed while filling the 5f sublevel. In this paper, we analyze some classes of compounds formed by the lanthanides with other elements of the Periodic Table, which can count into the thousands of binary compounds. The special place of lanthanides in the Periodic System of Elements made it possible to establish strict nonlinear relationships between the standard entropy and the lanthanide atomic number of the compounds Ln2X3 (X = O, S, Se, Te), LnN, LnB4, and LnF3 in the solid state. This relationship, based on tetrad-effect, can be applied to other physical and chemical properties of the isostructural compounds. The thermodynamic properties of actinides have been studied much less than lanthanides, but the similarity of physicochemical properties makes it possible for us to estimate, with sufficient accuracy, unexplored properties using fundamental laws. One of these laws is the tetrad-effect concept that is an effective tool to predict missing thermodynamic values for lanthanide and actinide compounds and to rationally plan experiments.
APA, Harvard, Vancouver, ISO, and other styles
9

Nielsen, Lea Gundorff, Anne Kathrine R. Junker, and Thomas Just Sørensen. "Composed in the f-block: solution structure and function of kinetically inert lanthanide(iii) complexes." Dalton Transactions 47, no. 31 (2018): 10360–76. http://dx.doi.org/10.1039/c8dt01501e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Wong-Ng, Winnie, Boris Paretzkin, and Edwin R. Fuller. "Crystal Chemistry and Phase Equilibria of the BaO-R2O3-CuO Systems." Advances in X-ray Analysis 33 (1989): 453–65. http://dx.doi.org/10.1154/s0376030800019881.

Full text
Abstract:
AbstractTwo important factors, the progressively decreasing size of the lanthanides, which is known as the lanthanide contraction, as well as the stability of different oxidation states of these elements influence the prediction of compound formation in the Ba-R-Cu-O systems. A systematic investigation of these lanthanide systems and comparison with the Y system has revealed a correlation of the effect of the above factors, in particular the size factor, on the trend of phase formation, solid solution formation and phase compatibility diagrams of the Ba-R-Cu-O systems. For example, it has been found that the smaller the size of R3+ or the greater the mismatch betwaen Ba2+ and R3+ in the solid solution series Ba2-ZR1+lCu3O6+x, the smaller the extent of solid solution formation. This differing extent of solid solution formation influences the ternary phase relationships.
APA, Harvard, Vancouver, ISO, and other styles
11

Onghena, Bieke, Eleonora Papagni, Ernesto Rezende Souza, Dipanjan Banerjee, Koen Binnemans, and Tom Vander Hoogerstraete. "Speciation of lanthanide ions in the organic phase after extraction from nitrate media by basic extractants." RSC Advances 8, no. 56 (2018): 32044–54. http://dx.doi.org/10.1039/c8ra06712k.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Gagné, Olivier Charles. "Bond-length distributions for ions bonded to oxygen: results for the lanthanides and actinides and discussion of the f-block contraction." Acta Crystallographica Section B Structural Science, Crystal Engineering and Materials 74, no. 1 (January 12, 2018): 49–62. http://dx.doi.org/10.1107/s2052520617017425.

Full text
Abstract:
Bond-length distributions have been examined for 84 configurations of the lanthanide ions and 22 configurations of the actinide ions bonded to oxygen, for 1317 coordination polyhedra and 10 700 bond distances for the lanthanide ions, and 671 coordination polyhedra and 4754 bond distances for the actinide ions. A linear correlation between mean bond length and coordination number is observed for the trivalent lanthanides ions bonded to O2−. The lanthanide contraction for the trivalent lanthanide ions bonded to O2− is shown to vary as a function of coordination number, and to diminish in scale with an increasing coordination number. The decrease in mean bond length from La3+ to Lu3+ is 0.25 Å for coordination number (CN) 6 (9.8%), 0.22 Å for CN 7 (8.7%), 0.21 Å for CN 8 (8.0%), 0.21 Å for CN 9 (8.2%) and 0.18 Å for CN 10 (6.9%). The crystal chemistry of Np5+ and Np6+ is shown to be very similar to that of U6+ when bonded to O2−, but differs for Np7+.
APA, Harvard, Vancouver, ISO, and other styles
13

Martín-Rodríguez, R., R. Valiente, F. Aguado, and A. C. Perdigón. "Highly efficient photoluminescence from isolated Eu3+ ions embedded in high-charge mica." J. Mater. Chem. C 5, no. 39 (2017): 10360–68. http://dx.doi.org/10.1039/c7tc01818e.

Full text
Abstract:
Incorporation of lanthanide ions in synthetic clay minerals is a promising approach to combine the efficient sharp-line emission of lanthanides with the unique structural stability and high adsorption capacity of high-charge micas.
APA, Harvard, Vancouver, ISO, and other styles
14

Liu, Jingjing, Lydia E. Nodaraki, Philip J. Cobb, Marcus J. Giansiracusa, Fabrizio Ortu, Floriana Tuna, and David P. Mills. "Synthesis and characterisation of light lanthanide bis-phospholyl borohydride complexes." Dalton Transactions 49, no. 19 (2020): 6504–11. http://dx.doi.org/10.1039/d0dt01241f.

Full text
Abstract:
Two families of lanthanide(iii) phospholyl borohydride complexes are reported (carbon = grey, hydrogen = white, oxygen = red, boron = yellow, phosphorus = magenta, potassium = blue, lanthanides = teal; only BH4 hydrogens are shown for clarity).
APA, Harvard, Vancouver, ISO, and other styles
15

Kovács, Eszter M., József Kónya, and Noémi M. Nagy. "Structural curiosities of lanthanide (Ln)-modified bentonites analyzed by radioanalytical methods." Journal of Radioanalytical and Nuclear Chemistry 322, no. 3 (September 19, 2019): 1747–54. http://dx.doi.org/10.1007/s10967-019-06765-6.

Full text
Abstract:
Abstract The effects of pH and lanthanide (La, Y) concentration were investigated on the release of iron from Ca-bentonite crystal structure. XRF results revealed that during the Ca–H cation exchange procedure iron loss was not observed. In the case of lanthanide modifications, the pH has low influence, meanwhile the concentration of lanthanide has high influence on iron loss. Thus, high amount of trivalent lanthanides cause the structural iron release.
APA, Harvard, Vancouver, ISO, and other styles
16

González Chávez, Fernando, and Hiram Isaac Beltrán. "Tuning dimensionality between 2D and 1D MOFs by lanthanide contraction and ligand-to-metal ratio." New Journal of Chemistry 45, no. 15 (2021): 6600–6610. http://dx.doi.org/10.1039/d0nj04055j.

Full text
Abstract:
2D/1D dimensionality tuning in LnMOFs is related to both (i) ligand-to-metal ratio and (ii) lanthanide contraction, this is only possible with Er/Tm, lighter lanthanides e.g. Pr only produced 2D MOFs, despite different ligand-to-metal ratios were used.
APA, Harvard, Vancouver, ISO, and other styles
17

Shahbazi, Shayan, C. J. Oldham, Austin D. Mullen, John D. Auxier II, and Howard L. Hall. "Synthesis, thermogravimetric analysis and enthalpy determination of lanthanide β-diketonates." Radiochimica Acta 107, no. 12 (November 26, 2019): 1173–84. http://dx.doi.org/10.1515/ract-2018-3085.

Full text
Abstract:
Abstract This work reports thermodynamic characterizations of lanthanide β-diketonates for use in nuclear fission product separation. Adsorption and sublimation enthalpies have been shown to be linearly correlated, therefore there is motivation to determine sublimation thermodynamics. An isothermal thermogravimetric analysis method is employed on fourteen lanthanide chelates for the ligands 2,2,6,6-tetramethyl-3,5-heptanedione and 6,6,7,7,8,8,8-heptafluoro-2,2-dimethyl-3,5-octanedione to determine sublimation enthalpies. No linear trend is seen across the series; values show a cyclical nature, possibly indicating a greater influence of chemisorption for some complexes and less of a role of physisorption in dictating adsorption differences between lanthanides in the same series. This is in line with previous reports in terms of the chromatographic separation order of the lanthanides. The results reported here can be used to manipulate separations parameters and column characteristics to better separate these lanthanide chelates. Fourteen chelates of the ligand 1,1,1-trifluoro-2,4-pentanedione are also thermally characterized but found to not sublime and be undesirable for this method. Additionally, all chelates are characterized by constant heating thermogravimetric analysis coupled with mass spectrometry, melting point analysis, elemental analysis and FTIR.
APA, Harvard, Vancouver, ISO, and other styles
18

Konchenko, Sergey N. "Reductive Approach to the Synthesis of the Molecular Lanthanide Polypnictide Complexes." Vestnik RFFI, no. 2 (June 25, 2019): 101–12. http://dx.doi.org/10.22204/2410-4639-2019-102-02-101-112.

Full text
Abstract:
In spite of the vigorous development during the last few decades, the coordination chemistry of lanthanides (Ln) is 95% still the chemistry of complexes with O- and N-donor ligands. The compounds with the Ln—E bond (E is a heavy element of the group 15th or 16th) up to now are considered as unconventional or exotic. Recently the fruitful “reductive approach” to this kind of compounds has been developed. The approach involves the performing of reactions between the strong reductants (Ln(II) complexes) and inorganic or organometallic compounds of the main group heavy elements. This paper is focused on the synthesis and structural diversity of the new family of the lanthanide compounds – polypnictide homo- and heterometallic complexes.
APA, Harvard, Vancouver, ISO, and other styles
19

Lincheneau, Christophe, Floriana Stomeo, Steve Comby, and Thorfinnur Gunnlaugsson. "Recent Highlights in the use of Lanthanide-directed Synthesis of Novel Supramolecular (Luminescent) Self-assembly Structures such as Coordination Bundles, Helicates and Sensors." Australian Journal of Chemistry 64, no. 10 (2011): 1315. http://dx.doi.org/10.1071/ch11184.

Full text
Abstract:
In this short review, we focus on the recent developments within the field of coordination chemistry where mono- or multimetallic supramolecular self-assemblies are formed by employing structurally defined organic ligands, taking advantage of the high coordination requirements of the lanthanides. Such synthesis results in the formation of both structurally complex and beautiful self-assemblies. Moreover, as the lanthanide ions possess both unique magnetic (e.g. GdIII and DyIII) and luminescent properties, either in the visible (EuIII, SmIII and TbIII) or near-infrared regions (YbIII, NdIII, ErIII), these physical features are usually transferred to the self-assemblies themselves, allowing the formation of highly functional structures, such as coordination networks, as well as molecular bundles and helicates. Hence, examples of the use of lanthanide-directed synthesis of luminescent sensors, some of which are formed on solid surfaces such as gold (flat surface or nanoparticles), and imaging agents are presented. Moreover, we demonstrate that by using chiral organic ligands, lanthanide-directed synthesis can also give rise to the formation of enantiomerically pure self-assemblies, the structure of which can be probed using circularly polarized luminescence.
APA, Harvard, Vancouver, ISO, and other styles
20

Zapp, Nicolas, and Holger Kohlmann. "The lanthanide hydride oxides SmHO and HoHO." Zeitschrift für Naturforschung B 73, no. 8 (August 28, 2018): 535–38. http://dx.doi.org/10.1515/znb-2018-0112.

Full text
Abstract:
AbstractMetal hydride oxides are an interesting class of compounds with potential for hydride ion conduction and as host materials for luminescence. SmHO and HoHO were prepared from mixtures of the sesquioxides Ln2O3 and the hydrides LnH2+x at 1173 K as gray powders (Ln=Sm, Ho). They crystallize in a fluorite type crystal structure with disordered anion distribution (Fm3̅m; SmHO: a=5.46953(6) Å, V=163.625(5) Å3; HoHO: a=5.27782(3) Å, V=147.016(2) Å3, based on powder X-ray diffraction) and show stability towards air. Lanthanide-oxygen and -hydrogen distances are 2.36838(3) Å in SmHO and 2.28536(1) Å in HoHO and comparable to those in binary lanthanide oxides and hydrides. Elemental analyses confirm the composition LnHO. Quantum-mechanical calculations show a negative enthalpy for the reaction RE2O3+REH3→3 REHO for all lanthanides and Y, with increasing values for decreasing ionic radii.
APA, Harvard, Vancouver, ISO, and other styles
21

Lobanov, N. N., and T. A. Kuznetsova. "Crystal chemistry of lanthanide oxochlorotungstates." Russian Journal of Inorganic Chemistry 53, no. 8 (August 2008): 1256–62. http://dx.doi.org/10.1134/s0036023608080184.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Kagan, Henri B. "Introduction: Frontiers in Lanthanide Chemistry." Chemical Reviews 102, no. 6 (June 2002): 1805–6. http://dx.doi.org/10.1021/cr020014i.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Evans, William J. "Perspectives in reductive lanthanide chemistry." Coordination Chemistry Reviews 206-207 (September 2000): 263–83. http://dx.doi.org/10.1016/s0010-8545(00)00267-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Zhu, Dao-Hong, Mary J. Kappel, and Kenneth N. Raymond. "Coordination chemistry of lanthanide catecholates." Inorganica Chimica Acta 147, no. 1 (July 1988): 115–21. http://dx.doi.org/10.1016/s0020-1693(00)80639-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Sieke, Corinna, Ingo Hartenbach, and Thomas Schleid. "Sulfidisch derivatisierte Oxodisilicate der schweren Lanthanide vom Formeltyp M4S3[Si2O7] (M = Gd - Tm) / Sulfide Derivatized Oxodisilicates of the Heavy Lanthanides with the Formula M4S3[Si2O7] (M = Gd - Tm)." Zeitschrift für Naturforschung B 57, no. 12 (December 1, 2002): 1427–32. http://dx.doi.org/10.1515/znb-2002-1214.

Full text
Abstract:
The sulfide derivatized oxodisilicates M4S3[Si2O7] of the heavy lanthanides (M = Gd - Tm) have been synthesized by reaction of M, M2O3, S, and SiO2 using MCl3 as flux in evacuated silica tubes at 850 °C for 7 d. The title compounds crystallize tetragonally in the space group I41/amd with 8 formula units per unit cell. Their crystal structure contains isolated [Si2O7]6− pyroanions of two vertex-sharing [SiO4] tetrahedra in eclipsed conformation with Si-O-Si bridging angles of about 130°. The two crystallographically independent lanthanide cations (M3+) both reside in tricapped trigonal prismatic environments of chalcogen ligands. In the case of M1 five S2− and three plus one O2− anions (CN = 8+1), for M2 three S2− and six O2− anions (CN = 9) fill the coordination spheres. Two of the three different sulfide anions are surrounded by four M3+ cations forming a seesaw (S2) and a square (S3), respectively. The third one (S1) is coordinated by six lanthanide cations in the form of a distorted octahedron. The whole structure is basically built up of cationic lanthanide silicate layers which are threaded by sulfide anions.
APA, Harvard, Vancouver, ISO, and other styles
26

Gao, Qi, Shuai Han, Qing Ye, Shuiyuan Cheng, Tianfang Kang, and Hongxing Dai. "Effects of Lanthanide Doping on the Catalytic Activity and Hydrothermal Stability of Cu-SAPO-18 for the Catalytic Removal of NOx (NH3-SCR) from Diesel Engines." Catalysts 10, no. 3 (March 17, 2020): 336. http://dx.doi.org/10.3390/catal10030336.

Full text
Abstract:
Lanthanide (La, Ce, Nd, Gd, Tb, Ho or Lu)-doped Cu-SAPO-18 samples were prepared using the ion-exchange method. Physicochemical properties of the samples were systematically characterized by a number of analytical techniques, and the effects of lanthanide doping on catalytic activity and hydrothermal stability of the Cu-SAPO-18 catalysts for the NH3-SCR reaction were examined. It is shown that the doping of lanthanide elements could affect the interaction between the active components (copper ions) and the AEI-structured SAPO-18 support. The inclusion of some lanthanides significantly slowed down hydrolysis of the catalyst during hydrothermal aging treatment process, leading to an enhanced catalytic activity at both low and high temperatures and hydrothermal stability. In particular, Ce doping promoted the Cu2+ ions to migrate to the energetically favorable sites for enhancement in catalytic activity, whereas the other lanthanide ions exerted little or an opposite effect on the migration of Cu2+ ions. Additionally, Ce doping could improve hydrothermal stability of the Cu-SAPO-18 catalyst by weakening hydrolysis of the catalyst during the hydrothermal aging treatment process. Ce doping increased the catalytic activity of Cu-SAPO-18 at low and high temperatures, which was attributed to modifications of the redox and/or isolated Cu2+ active centers.
APA, Harvard, Vancouver, ISO, and other styles
27

NAKASHIMA, Nobuaki. "Unconventional Laser Chemistry. Laser Chemistry of Lanthanide Ions." Review of Laser Engineering 24, no. 7 (1996): 787–95. http://dx.doi.org/10.2184/lsj.24.787.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Flakina, Alexandra M., Elena I. Zhilyaeva, Gennady V. Shilov, Maxim A. Faraonov, Svetlana A. Torunova, and Dmitri V. Konarev. "Layered Organic Conductors Based on BEDT-TTF and Ho, Dy, Tb Chlorides." Magnetochemistry 8, no. 11 (October 28, 2022): 142. http://dx.doi.org/10.3390/magnetochemistry8110142.

Full text
Abstract:
Molecular semiconductors with lanthanide ions have been synthesized based on BEDT-TTF and lanthanide chlorides: (BEDT-TTF)2[HoCl2(H2O)6]Cl2(H2O)2 (1, which contains a 4f holmium cation), and (BEDT-TTF)2LnCl4(H2O)n (Ln = Dy, Tb, Ho (2–4), which contain 4f anions of lanthanides). Conductivity and EPR measurements have been carried out along with the SQUID magnetometry, and the crystal structure has been established for 1. The structure of 1 is characterized by an alternation of organic radical cation layers composed of BEDT-TTF chains and inorganic layers consisting of chains of the [HoCl2(H2O)6]+ cations interlinked by chlorine anions and crystallization water molecules. The magnetic susceptibility of 1–3 determined mainly by lanthanide ions follows the Curie–Weiss law with the Weiss temperatures of −3, −3, −2 K for 1–3, respectively, indicating weak antiferromagnetic coupling between paramagnetic lanthanide ions. The signals attributed to the BEDT-TTF+· radical cations only are observed in the EPR spectra of 1–3, which makes it possible to study their magnetic behavior. There are two types of chains in the organic layers of 1: the chains composed of neutral molecules and those formed by BEDT-TTF+· radical cations. As a result, uniform 1D antiferromagnetic coupling of spins is observed in the BEDT-TTF+· chains with estimated exchange interaction J = −10 K. The study of dynamic magnetic properties of 1–3 shows that these compounds are not SMMs.
APA, Harvard, Vancouver, ISO, and other styles
29

Alakhras, Fadi. "Kinetic Studies on the Removal of Some Lanthanide Ions from Aqueous Solutions Using Amidoxime-Hydroxamic Acid Polymer." Journal of Analytical Methods in Chemistry 2018 (July 8, 2018): 1–7. http://dx.doi.org/10.1155/2018/4058503.

Full text
Abstract:
Lanthanide metal ions make distinctive and essential contributions to recent global proficiency. Extraction and reuse of these ions is of immense significance especially when the supply is restricted. In light of sorption technology, poly(amidoxime-hydroxamic) acid sorbents are synthesized and utilized for the removal of various lanthanide ions (La3+, Nd3+, Sm3+, Gd3+, and Tb3+) from aqueous solutions. The sorption speed of trivalent lanthanides (Ln3+) depending on the contact period is studied by a batch equilibrium method. The results reveal fast rates of metal ion uptake with highest percentage being achieved after 15–30 min. The interaction of poly(amidoxime-hydroxamic) acid sorbent with Ln3+ ions follows the pseudo-second-order kinetic model with a correlation coefficient R2 extremely high and close to unity. Intraparticle diffusion data provide three linear plots indicating that the sorption process is affected by two or more steps, and the intraparticle diffusion rate constants are raised among reduction of ionic radius of the studied lanthanides.
APA, Harvard, Vancouver, ISO, and other styles
30

Pallares, Roger M., and Rebecca J. Abergel. "Transforming lanthanide and actinide chemistry with nanoparticles." Nanoscale 12, no. 3 (2020): 1339–48. http://dx.doi.org/10.1039/c9nr09175k.

Full text
Abstract:
This minireview summarizes and discusses recent progress on the use of nanoparticles in lanthanide and actinide chemistry. We examine different types of nanoparticles and critically analyze their performance in a comparative mode.
APA, Harvard, Vancouver, ISO, and other styles
31

Bourne, S. A., and L. J. Moitsheki. "Physical chemistry of lanthanide coordination polymers." Acta Crystallographica Section A Foundations of Crystallography 60, a1 (August 26, 2004): s98. http://dx.doi.org/10.1107/s0108767304098071.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Parker, David, and J. A. Gareth Williams. "Getting excited about lanthanide complexation chemistry." Journal of the Chemical Society, Dalton Transactions, no. 18 (1996): 3613. http://dx.doi.org/10.1039/dt9960003613.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Arnaud-Neu, F. "Solution chemistry of lanthanide macrocyclic complexes." Chemical Society Reviews 23, no. 4 (1994): 235. http://dx.doi.org/10.1039/cs9942300235.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Payne, R. F., S. M. Schulte, M. Douglas, J. I. Friese, O. T. Farmer, and E. C. Finn. "Investigation of gravity lanthanide separation chemistry." Journal of Radioanalytical and Nuclear Chemistry 287, no. 3 (October 26, 2010): 863–67. http://dx.doi.org/10.1007/s10967-010-0838-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Roesky, Peter W. "P?N ligands in lanthanide chemistry." Heteroatom Chemistry 13, no. 6 (2002): 514–20. http://dx.doi.org/10.1002/hc.10096.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Ferru, Geoffroy, Benjamin Reinhart, Mrinal K. Bera, Monica Olvera de la Cruz, Baofu Qiao, and Ross J. Ellis. "The Lanthanide Contraction beyond Coordination Chemistry." Chemistry - A European Journal 22, no. 20 (April 6, 2016): 6899–904. http://dx.doi.org/10.1002/chem.201601032.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Akitsu, Takashiro. "Lanthanide Complexes in Recent Molecules." Molecules 27, no. 18 (September 15, 2022): 6019. http://dx.doi.org/10.3390/molecules27186019.

Full text
Abstract:
The MDPI journal Molecules publishes excellent papers on molecules in every issue, but I found that there are surprisingly only a few papers (3) and reviews (1) on rare-earth elements (lanthanide) in the category of inorganic chemistry these days [...]
APA, Harvard, Vancouver, ISO, and other styles
38

Southworth, Faye Y., Claire Wilson, Simon J. Coles, and Andrew M. Fogg. "Synthesis and characterisation of a new anion exchangeable layered hydroxyiodide." Dalton Trans. 43, no. 27 (2014): 10451–55. http://dx.doi.org/10.1039/c4dt00123k.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Kazeminejad, Neda, Denise Munzel, Michael T. Gamer, and Peter W. Roesky. "Bis(amidinate) ligands in early lanthanide chemistry – synthesis, structures, and hydroamination catalysis." Chemical Communications 53, no. 6 (2017): 1060–63. http://dx.doi.org/10.1039/c6cc08958e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Todorovsky, Dimitr S., Miroslava M. Getsova, Maria M. Milanova, Masato Kakihana, Nikolina L. Petrova, Michail G. Arnaudov, and Venelin G. Enchev. "The chemistry of the processes involved in the production of lanthanide titanates by the polymerized-complex method." Canadian Journal of Chemistry 85, no. 7-8 (July 1, 2007): 547–59. http://dx.doi.org/10.1139/v07-067.

Full text
Abstract:
The composition, some spectral characteristics, and thermal decomposition of solid lanthanide–titanium (lanthanide (Ln) = Y, La, Ce) and lanthanide–titanium citrates (CA) and tartrates (TA) have been studied. The complexes have been prepared in ethylene glycol medium at conditions modeling those of the polymerized-complex method applied for Ln2Ti2O7 preparation. Special attention has been paid to the chemical nature of the bimetallic products as well as to the factors influencing the deprotonation of the alcoholic OH groups of the acidic ligands. The results contribute to further elucidation of the complexation and thermal decomposition processes involved in the polymerized-complex method.Key words: inorganic compounds, sol–gel chemistry, infrared spectroscopy, nuclear magnetic resonance, thermogravimetric analysis (TGA).
APA, Harvard, Vancouver, ISO, and other styles
41

Daszczyńska, Agnieszka, Tomasz Krucoń, Robert Stasiuk, Marta Koblowska, and Renata Matlakowska. "Lanthanide-Dependent Methanol Metabolism of a Proteobacteria-Dominated Community in a Light Lanthanide-Rich Deep Environment." International Journal of Molecular Sciences 23, no. 7 (April 1, 2022): 3947. http://dx.doi.org/10.3390/ijms23073947.

Full text
Abstract:
This study investigated the occurrence and diversity of proteobacterial XoxF-type methanol dehydrogenases (MDHs) in the microbial community that inhabits a fossil organic matter- and sedimentary lanthanide (Ln3+)-rich underground mine environment using a metagenomic and metaproteomic approach. A total of 8 XoxF-encoding genes (XoxF-EGs) and 14 protein sequences matching XoxF were identified. XoxF-type MDHs were produced by Alpha-, Beta-, and Gammaproteobacteria represented by the four orders Methylococcales, Nitrosomonadales, Rhizobiales, and Xanthomonadales. The highest number of XoxF-EG- and XoxF-matching protein sequences were affiliated with Nitrosomonadales and Rhizobiales, respectively. Among the identified XoxF-EGs, two belonged to the XoxF1 clade, five to the XoxF4 clade, and one to the XoxF5 clade, while seven of the identified XoxF proteins belonged to the XoxF1 clade, four to the XoxF4 clade, and three to the XoxF5 clade. Moreover, the accumulation of light lanthanides and the presence of methanol in the microbial mat were confirmed. This study is the first to show the occurrence of XoxF in the metagenome and metaproteome of a deep microbial community colonizing a fossil organic matter- and light lanthanide-rich sedimentary environment. The presented results broaden our knowledge of the ecology of XoxF-producing bacteria as well as of the distribution and diversity of these enzymes in the natural environment.
APA, Harvard, Vancouver, ISO, and other styles
42

Bennett, Stacey D., Bryony A. Core, Matthew P. Blake, Simon J. A. Pope, Philip Mountford, and Benjamin D. Ward. "Chiral lanthanide complexes: coordination chemistry, spectroscopy, and catalysis." Dalton Trans. 43, no. 15 (2014): 5871–85. http://dx.doi.org/10.1039/c4dt00114a.

Full text
Abstract:
Luminescent lanthanide complexes bearing amido-bisoxazoline ligands are reported. They were studied using time-resolved luminescence spectroscopy, and were probed for their activity in hydroamination/cyclisation and ring-opening polymerisation catalysis.
APA, Harvard, Vancouver, ISO, and other styles
43

Berwick, Matthew R., Louise N. Slope, Caitlin F. Smith, Siobhan M. King, Sarah L. Newton, Richard B. Gillis, Gary G. Adams, et al. "Location dependent coordination chemistry and MRI relaxivity, in de novo designed lanthanide coiled coils." Chemical Science 7, no. 3 (2016): 2207–16. http://dx.doi.org/10.1039/c5sc04101e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Wang, S. G., and W. H. E. Schwarz. "Lanthanide Diatomics and Lanthanide Contractions." Journal of Physical Chemistry 99, no. 30 (July 1995): 11687–95. http://dx.doi.org/10.1021/j100030a011.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Zucchi, Gaël. "The Utility of 2,2′-Bipyrimidine in Lanthanide Chemistry: From Materials Synthesis to Structural and Physical Properties." International Journal of Inorganic Chemistry 2011 (June 2, 2011): 1–13. http://dx.doi.org/10.1155/2011/918435.

Full text
Abstract:
This paper reviews the recent investigations undertaken on the use of 2,2′-bipyrimidine (bpm) as a ligand for designing molecular complexes as well as polymeric lanthanide materials. A special emphasis is put on the ability of this polydentate neutral ligand to yield compounds of various dimensionalities, to act as a connector between these large ions, and influence their emissive and magnetic properties. This ligand can adopt a terminal or a bridging coordination mode with lanthanide ions, thus generating a wealth of frameworks of various topologies with the 4f elements. The main focus of this review is to show the originality brought by bpm in lanthanide structural chemistry and solid-state photophysics and magnetism.
APA, Harvard, Vancouver, ISO, and other styles
46

Fang, Xin, Li-Mao Cai, Yu-Chen Shao, and Mei-Jin Lin. "Lanthanide contraction in linear lanthanide–oxygen clusters." Journal of Coordination Chemistry 67, no. 21 (October 14, 2014): 3542–50. http://dx.doi.org/10.1080/00958972.2014.969723.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Martinez-Martin, Paloma, Josefina Perles, and Juan Carlos Rodriguez-Ubis. "Crystal Structure Dependence of the Energy Transfer from Tb(III) to Yb(III) in Metal–Organic Frameworks Based in Bispyrazolylpyridines." Crystals 10, no. 2 (January 27, 2020): 69. http://dx.doi.org/10.3390/cryst10020069.

Full text
Abstract:
Luminescent mixed lanthanide metal−organic framwork (MOF) materials have been prepared from two polyheterocyclic diacid ligands, 2,6-bis(3-carboxy-1-pyrazolyl)pyridine and 2,6-bis(4-carboxy-1-pyrazolyl)pyridine. The crystal structures of the two organic molecules are presented together with the structures for the MOFs obtained by hydrothermal synthesis either with Yb(III) or mixed Tb(III)/Yb(III) ions. Different coordination architectures result from each ligand, revealing also important differences between the lanthanides. The mixed lanthanide metal−organic frameworks also present diverse luminescent behavior; in the case of 2,6-bis(4-carboxy-1-pyrazolyl)pyridine, where no coordinated water is present in the metal environment, Tb(III) and Yb(III) characteristic emission is observed by excitation of the bispyrazolylpyridine chromophore.
APA, Harvard, Vancouver, ISO, and other styles
48

Nicholas, Hannah M., Michele Vonci, Conrad A. P. Goodwin, Song Wei Loo, Siobhan R. Murphy, Daniel Cassim, Richard E. P. Winpenny, Eric J. L. McInnes, Nicholas F. Chilton, and David P. Mills. "Electronic structures of bent lanthanide(III) complexes with two N-donor ligands." Chemical Science 10, no. 45 (2019): 10493–502. http://dx.doi.org/10.1039/c9sc03431e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Broker, Grant A., Marc A. Klingshirn, and Robin D. Rogers. "Green chemistry and lanthanide-based crystal engineering." Journal of Alloys and Compounds 344, no. 1-2 (October 2002): 123–27. http://dx.doi.org/10.1016/s0925-8388(02)00349-3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Aspinall, Helen C. "Chiral Lanthanide Complexes: Coordination Chemistry and Applications." Chemical Reviews 102, no. 6 (June 2002): 1807–50. http://dx.doi.org/10.1021/cr010288q.

Full text
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