[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (298)

Search Parameters:
Keywords = lanthanide complex

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 6094 KiB  
Article
A Fluorine-Functionalized Tb(III)–Organic Framework for Ba2+ Detection
by Yang Zhang, Hua Tan, Jiaping Zhu, Linhai Duan, Yuchi Ding, Fenglan Liang, Yongshi Li, Xinteng Peng, Ruomei Jiang, Jiaxin Yu, Jianjiong Fan, Yuhang Chen, Rimeng Chen and Deyun Ma
Molecules 2024, 29(24), 5903; https://doi.org/10.3390/molecules29245903 - 13 Dec 2024
Viewed by 469
Abstract
The development of lanthanide–organic frameworks (Ln-MOFs) using for luminescence sensing and selective gas adsorption applications is of great significance from an energy and environmental perspective. This study reports the solvothermal synthesis of a fluorine-functionalized 3D microporous Tb-MOF with a face-centered cubic (fcu [...] Read more.
The development of lanthanide–organic frameworks (Ln-MOFs) using for luminescence sensing and selective gas adsorption applications is of great significance from an energy and environmental perspective. This study reports the solvothermal synthesis of a fluorine-functionalized 3D microporous Tb-MOF with a face-centered cubic (fcu) topology constructed from hexanuclear clusters (Tb6O30) bridged by fdpdc ligands, formulated as {[Tb6(fdpdc)6(μ3-OH)8(H2O)6]·4DMF}n (1), (fdpdc = 3-fluorobiphenyl-4,4′-dicarboxylate). Complex 1 displays a 3D framework with the channel of 7.2 × 7.2 Å2 (measured between opposite atoms) perpendicular to the a-axis. With respect to Ba2+ cation, the framework of activated 1 (1a) exhibits high selectivity and reversibility in luminescence sensing function, with an LOD of 4.34665 mM. According to the results of simulations, compared to other small gas molecules (CO2, N2, H2, CO, and CH4), activated 1 (1a) shows a high adsorption selectivity for C2H2 at 298 K. Full article
(This article belongs to the Special Issue Inorganic Chemistry in Asia)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Single-capped tetragonal prism of Tb(III) in <b>1</b>, blue line, configuration bar; red bullet, oxygen atoms; green circle, Tb. The coordination environment of Zn(II) ions in <b>1</b>. (<b>b</b>) Hexanuclear Tb<sub>6</sub>O<sub>30</sub> unit of <b>1</b>. (<b>c</b>) View of the 3D framework packing perpendicular to the <span class="html-italic">a</span>-axis. (<b>d</b>) The <b>fcu</b> topology network of <b>1</b>.</p>
Full article ">Figure 2
<p>PXRD patterns of <b>1</b>.</p>
Full article ">Figure 3
<p>(<b>a</b>) The fluorescence excitation and emission of <b>1</b> and H<sub>2</sub>fdpdc ligand in solid state. (<b>b</b>) The cation-dependent fluorescence emission of <b>1</b>. (<b>c</b>) The luminescence intensity at 545 nm with the presence of different cations. (<b>d</b>) The luminescent emission spectra of <b>1</b> in Ba<sup>2+</sup> solutions with different concentrations.</p>
Full article ">Figure 4
<p>(<b>a</b>) The correlation between relative intensity and the addition of Ba<sup>2+</sup> ion at different concentrations over time at 545 nm. (<b>b</b>) Calculated LOD of <b>1</b> in Ba<sup>2+</sup> solutions.</p>
Full article ">Figure 5
<p>(<b>a</b>) Luminescence intensity of <b>1</b> at 545 nm dispersed in water with the addition of different metal ions. The concentration of interferential metal ions and Ba<sup>2+</sup> ion are both 1 mM. (<b>b</b>) Luminescence intensity (545 nm) of <b>1</b> during five recycling.</p>
Full article ">Figure 6
<p>(<b>a</b>) N<sub>2</sub> sorption isotherm of <b>1a</b> at 77 K (the inset shows the pore size distribution). (<b>b</b>) Room temperature adsorption isotherms of C<sub>2</sub>H<sub>2</sub>, CO<sub>2</sub>, N<sub>2</sub>, H<sub>2</sub>, CO, and CH<sub>4</sub> based on the simulation. (<b>c</b>) Isosteric heat of adsorption of C<sub>2</sub>H<sub>2</sub> for <b>1</b> calculated based on molecular simulation isotherms at 298 K. (<b>d</b>) Selectivities for C<sub>2</sub>H<sub>2</sub> in the equimolar C<sub>2</sub>H<sub>2</sub>/CO<sub>2</sub>, C<sub>2</sub>H<sub>2</sub>/N<sub>2</sub>, C<sub>2</sub>H<sub>2</sub>/H<sub>2</sub>, C<sub>2</sub>H<sub>2</sub>/CO, and C<sub>2</sub>H<sub>2</sub>/CH<sub>4</sub> at 298 K from the simulation.</p>
Full article ">
20 pages, 6567 KiB  
Article
Calixarene-like Lanthanide Single-Ion Magnets Based on NdIII, GdIII, TbIII and DyIII Oxamato Complexes
by Tamyris T. da Cunha, João Honorato de Araujo-Neto, Meiry E. Alvarenga, Felipe Terra Martins, Emerson F. Pedroso, Davor L. Mariano, Wallace C. Nunes, Nicolás Moliner, Francesc Lloret, Miguel Julve and Cynthia L. M. Pereira
Magnetochemistry 2024, 10(12), 103; https://doi.org/10.3390/magnetochemistry10120103 - 12 Dec 2024
Viewed by 540
Abstract
In this work, we describe the synthesis, crystal structures and magnetic properties of four air-stable mononuclear lanthanide(III) complexes with the N-(2,4,6-trimethylphenyl)oxamate (Htmpa) of formula: n-Bu4N[Nd(Htmpa)4(H2O)]·4H2O (1), n-Bu4N[Gd(Htmpa)4 [...] Read more.
In this work, we describe the synthesis, crystal structures and magnetic properties of four air-stable mononuclear lanthanide(III) complexes with the N-(2,4,6-trimethylphenyl)oxamate (Htmpa) of formula: n-Bu4N[Nd(Htmpa)4(H2O)]·4H2O (1), n-Bu4N[Gd(Htmpa)4(H2O)]·3DMSO·2H2O (2), n-Bu4N[Tb(Htmpa)4(H2O)]·3DMSO·1H2O (3) and n-Bu4N[Dy(Htmpa)4(H2O)]·3DMSO·2H2O (4) (n-Bu4N+ = n-tetrabutylammonium; DMSO = dimethylsulfoxide). Their crystal structures reveal the occurrence of calixarene-type monoanionic species containing all-cis-disposed Htmpa ligands and one water molecule coordinated with the respective LnIII ion (Ln = Nd, Gd, Tb and Dy), featuring a nine-coordinated environment with muffin (MFF-9) (1) or spherical-capped square antiprism (CSAPR-9) (24) geometry. The major difference between their crystal structures is related to the nature of crystallization solvent molecules, either water (1) or both DMSO and water (24). The intermolecular hydrogen bonds among the self-complementary Htmpa ligands in all four compounds mediated a 2 D supramolecular network in the solid state. Direct-current (dc) magnetic properties for 14 show typical behavior for the ground state terms of the LnIII ions [4I9/2 (Nd); 8S7/2(Gd), 7F6 (Tb), 6H15/2 (Dy)]. Alternating-current (ac) magnetic measurements reveal the presence of slow magnetic relaxation without the presence of a dc field only for 4. In contrast, field-induced slow magnetic relaxation behavior was found in complexes 1, 2 and 3. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>ORTEP view of the crystal structures of (<b>a</b>) <b>1</b> and (<b>b</b>) <b>3</b>. Crystallization water molecules and hydrogen atoms have been omitted for clarity, and ellipsoids represent 50% probability levels.</p>
Full article ">Figure 2
<p>(<b>a</b>) Top and side views of the coordination polyhedron around the Nd1 metal center of <b>1</b> with the atom numbering scheme for the donor atoms, showing the distorted muffin geometry. The O1D/O3C/O3D atoms compose the triangular base, and the O1A/O1C/O3A/O3B/O1w atoms compose the pentagonal plane with the O1B atom in the capped position. (<b>b</b>) Top and side views of the coordination polyhedron around the Tb1 metal center of <b>3</b> with the atom numbering scheme for the donor atoms, showing the distorted capped square antiprism geometry. The O3A/O3B/O3C/O3D atoms compose the square base of the prism, and the O1A/O1B/O1C/O1D atoms compose the quadratic plane of the capped face with the O1W atom in the capped position.</p>
Full article ">Figure 3
<p>(<b>a</b>) View of the supramolecular layers of hydrogen-bonded mononuclear units of <b>1</b> along the crystallographic <span class="html-italic">ab</span> plane. (<b>b</b>) View of the supramolecular layers of hydrogen-bonded mononuclear units of <b>3</b> along the crystallographic <span class="html-italic">ac</span> plane, showing the presence of additional hydrogen-bonded DMSO molecules. Cyan dashed lines represent hydrogen bonds. CH hydrogen atoms, remaining solvent molecules and tetrabutylammonium cations were omitted for the sake of clarity.</p>
Full article ">Figure 4
<p>Views perpendicular to the layers growing onto the (<b>a</b>) <span class="html-italic">ab</span> plane in <b>1</b> and (<b>b</b>) <span class="html-italic">ac</span> plane in <b>3</b>. Three layers are depicted in each panel, while hydrogen atoms and solvent molecules were omitted for the sake of clarity. Color codes: carbon, gray; nitrogen, light blue; oxygen, red; Nd, light green; and terbium, green.</p>
Full article ">Figure 5
<p><span class="html-italic">χ</span><sub>M</sub><span class="html-italic">T</span> vs. <span class="html-italic">T</span> curves for <b>1</b>–<b>4</b> (<b>a</b>–<b>d</b>). Inset: <span class="html-italic">M</span> vs. <span class="html-italic">H</span> curves. Solid lines represent the best-fit curves according to the text.</p>
Full article ">Figure 6
<p>(<b>a</b>) Representation of the easy magnetization axis of dysprosium(III) complex <b>4</b> as a red arrow in each single-ion magnet (color code: Dy in violet, N in blue, oxygen in red, carbon in grey). (<b>b</b>) The easy magnetization axis is represented by red arrows in the crystal packing of <b>4</b> along the crystallographic <span class="html-italic">ac</span> direction. For clarity, the hydrogen atoms, solvent molecules and counterions were omitted.</p>
Full article ">Figure 7
<p>(<b>a</b>) <span class="html-italic">χ</span><sub>M</sub>′ and <span class="html-italic">χ</span><sub>M</sub>″ vs. <span class="html-italic">ν</span> curves of <b>1</b> under 1.0 kOe dc magnetic field. (<b>b</b>) Arrhenius plot of <b>1</b> under 1.0 kOe dc magnetic field. Solid lines represent fits to the data (see text).</p>
Full article ">Figure 8
<p>(<b>a</b>) <span class="html-italic">χ</span><sub>M</sub>′ and <span class="html-italic">χ</span><sub>M</sub>″ vs<span class="html-italic">. ν</span> curves of <b>2</b> under 1.0 kOe dc magnetic field. (<b>b</b>) Arrhenius plot of <b>2</b> under a 1.0 kOe dc magnetic field. Solid lines represent fits to the data (see text).</p>
Full article ">Figure 9
<p>(<b>a</b>) <span class="html-italic">χ</span><sub>M</sub>′ and <span class="html-italic">χ</span><sub>M</sub>″ vs. <span class="html-italic">ν</span> curves of <b>3</b> under 5.0 kOe dc magnetic field. (<b>b</b>) Arrhenius plot of <b>3</b> under 5.0 kOe dc magnetic field. Solid lines represent fits to the data (see text).</p>
Full article ">Figure 10
<p><span class="html-italic">χ</span><sub>M</sub>′ and <span class="html-italic">χ</span><sub>M</sub>″ vs. <span class="html-italic">ν</span> curves of compound <b>4</b> under (<b>a</b>) zero dc magnetic field and (<b>b</b>) 1.0 kOe dc magnetic field. Arrhenius plots of <b>4</b> (<b>c</b>) under zero dc field and (<b>d</b>) 1.0 kOe dc field. Solid lines represent fits to the data (see text).</p>
Full article ">Scheme 1
<p>(<b>a</b>) Chemical structure of the proligand <span class="html-italic">N</span>-(2,4,6-trimethylphenyl)oxamic acid ethyl ester (EtHtmpa) and (<b>b</b>) their mononuclear lanthanide(III) oxamate complexes.</p>
Full article ">Scheme 2
<p>Synthetic procedure for complexes <b>1</b>–<b>4</b>.</p>
Full article ">
12 pages, 1481 KiB  
Article
Thiophenyl Anilato-Based NIR-Emitting Lanthanide (LnIII = Er, Yb) Dinuclear Complexes
by Fabio Manna, Mariangela Oggianu, Valentina Mameli, Stefano Lai, Angelica Simbula, Francesco Quochi, Narcis Avarvari and Maria Laura Mercuri
Molecules 2024, 29(23), 5804; https://doi.org/10.3390/molecules29235804 - 9 Dec 2024
Viewed by 479
Abstract
By combining ErIII and YbIII ions with 3,6-dithiophene-anilate (Th2An) and scorpionate hydrotris(pyrazol-1-yl)borate (HBpz3) ligands new luminescent dinuclear complexes are obtained. The two materials formulated as [((HB(pz)3)2Yb)2(μ-th2An)]·4DCM·1.3H2O [...] Read more.
By combining ErIII and YbIII ions with 3,6-dithiophene-anilate (Th2An) and scorpionate hydrotris(pyrazol-1-yl)borate (HBpz3) ligands new luminescent dinuclear complexes are obtained. The two materials formulated as [((HB(pz)3)2Yb)2(μ-th2An)]·4DCM·1.3H2O 1Yb and [((HB(pz)3)2Er)2(μ-th2An)]·4DCM·1.8H2O 1Er, respectively, have been structurally characterized by SC-XRD and PXRD studies. This study presents a comprehensive investigation of the photophysical properties of the Th2An ligand for the first time. Our findings reveal the crucial role of the thiophene anilate as an effective optical antenna, which sensitizes near-infrared (NIR)-emitting lanthanide ions, specifically ErIII and YbIII. The significant impact of vibrational quenching on the LnIII NIR emission efficiency has been also highlighted. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structural representations of the dinuclear complexes <b>1Yb</b> (<b>a</b>) and <b>1Er</b> (<b>b</b>) with thermal ellipsoids drawn at a 50% probability level. Hydrogen atoms and co-crystallized DCM molecules have been omitted for clarity. Color code: C (grey), N (blue), O (red), S (yellow), B (pink), Yb (green) and Er (light green).</p>
Full article ">Figure 2
<p>Diffuse reflectance and emission spectra of Ln<sup>III</sup> dinuclear complexes in the solid state. (<b>a</b>) UV-Vis range, showing ligand-centered absorption and emission (excitation at 350 nm). (<b>b</b>) NIR range, highlighting Ln<sup>III</sup>-centered emission (excited at 355 nm) and vibrational absorption bands. Spectroscopic terms for the Ln<sup>III</sup> upper manifolds, as well as key vibrational overtones and combination modes, are labeled. Diffuse reflectance spectra are represented by solid lines, while PL spectra are shown as dotted lines in both panels.</p>
Full article ">Figure 3
<p>Time-resolved NIR PL of Ln<sup>III</sup> dinuclear complexes in the solid state (excitation at 355 nm). Inset: Time-resolved and spectrally integrated Vis PL of Ln<sup>III</sup> complexes in the solid state (excitation at 350 nm). Dots: Experimental data; Black lines: Biexponential decay fits; refer to the text for details.</p>
Full article ">Scheme 1
<p>Synthesis of complexes <b>1Yb</b> and <b>1Er</b>.</p>
Full article ">
10 pages, 2942 KiB  
Article
Synthesis, Structure, and Properties of 2D Lanthanide(III) Coordination Polymers Constructed from Cyclotriphosphazene-Functionlized Hexacarboxylate Ligand
by Qi Jia, Yicheng Yao, Xiaoming Zhu, Juntao Wang, Zeyu Li, Liudi Ji and Peng Hu
Molecules 2024, 29(23), 5602; https://doi.org/10.3390/molecules29235602 - 27 Nov 2024
Viewed by 450
Abstract
The design and synthesis of novel lanthanide-based coordination polymers (Ln-CPs) from flexible organic ligands is still attractive and challenging. In this work, two isostructural Ln-CPs with a unique 2D network, namely, [Ln2(H3L)2(DMF)]]n (Ln = Dy for [...] Read more.
The design and synthesis of novel lanthanide-based coordination polymers (Ln-CPs) from flexible organic ligands is still attractive and challenging. In this work, two isostructural Ln-CPs with a unique 2D network, namely, [Ln2(H3L)2(DMF)]]n (Ln = Dy for 1, Tb for 2) based on a flexible polycarboxylic acid ligand hexakis(4-carboxylato-phenoxy)cyclotriphosphazene (H6L), have been solvothermally synthesized and structurally characterized. Significantly, it is the first observation of polycarboxylic acid ligands participating in coordination in the construction of coordination polymers in the form of semi-deprotonation. Magnetic measurements showed the presence of field-induced slow magnetic relaxation in complex 1. The luminescence property of 2 had been studied in the solid state at room temperature. Full article
(This article belongs to the Special Issue Advances in Coordination Chemistry 2.0)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The perspective view of the asymmetric unit of compound <b>1</b> with partial atoms labeled. Hydrogen atoms have been omitted for clarity. (<b>b</b>) The coordination geometry of the Dy1 ion (<b>left</b>) and Dy2 (<b>right</b>) ion in compound <b>1</b>. Color code: Dy<sup>III</sup>, bright green; O, red; N, blue; P, plum; C, gray.</p>
Full article ">Figure 2
<p>The connection mode of the semi-protonated hexacarboxylic acid ligand (<b>a</b>) and dinuclear unit (<b>b</b>) of compound <b>1</b>. Asymmetry code: A, x, 1/2 − y, −1/2 + z; B, 1 − x, −1/2 + y, 3/2 − z. Color code: Dy<sup>III</sup>, bright green; O, red; N, blue; P, plum; C, gray; H, sky blue.</p>
Full article ">Figure 3
<p>The 2D network structure of compound <b>1</b> with ten-membered metal rings along the a-axis.</p>
Full article ">Figure 4
<p>(<b>a</b>) Plots of <span class="html-italic">χ</span><sub>M</sub><span class="html-italic">T</span> vs. <span class="html-italic">T</span> for compound <b>1</b>. (<b>b</b>) <span class="html-italic">M</span> vs. <span class="html-italic">H</span> and <span class="html-italic">M</span> vs. <span class="html-italic">H</span>/<span class="html-italic">T</span> (inset) plots for compound <b>1</b> at 2–5 K. The solid lines show a guide for the eyes.</p>
Full article ">Figure 5
<p>Temperature dependence of the in-phase and out-of-phase ac magnetic susceptibility signals at the indicated frequencies for compound <b>1</b> under the optimal dc field.</p>
Full article ">Figure 6
<p>Room temperature solid-state emission spectra of compound <b>2</b> and H<sub>6</sub>L under excitation of 352 nm.</p>
Full article ">Scheme 1
<p>The structure of hexacarboxylic acid ligand H<sub>6</sub>L used in this work.</p>
Full article ">
27 pages, 6065 KiB  
Article
Extraction of Lanthanides(III) from Nitric Acid Solutions with N,N′-dimethyl-N,N′-dicyclohexyldiglycolamide into Bis(trifluoromethylsulfonyl)imide-Based Ionic Liquids and Their Mixtures with Molecular Organic Diluents
by Alexander N. Turanov, Vasilii K. Karandashev, Vladimir E. Baulin, Yury M. Shulga and Dmitriy V. Baulin
Minerals 2024, 14(11), 1167; https://doi.org/10.3390/min14111167 - 17 Nov 2024
Viewed by 662
Abstract
The extraction of lanthanides(III) from aqueous nitric acid solutions with novel unsymmetrical diglycolamide extactant, N,N′-dimethyl-N,N′-dicyclohexyldiglycolamide (DMDCHDGA) into bis(trifluoromethylsulfoyl)imide-based ionic liquids (ILs), namely 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([C4mim][Tf2N]), 1-octyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([C8mim][Tf2N]), benzyltriethylammonium bis(trifluoromethylsulfonyl)imide ([N222Bn][Tf2N]) [...] Read more.
The extraction of lanthanides(III) from aqueous nitric acid solutions with novel unsymmetrical diglycolamide extactant, N,N′-dimethyl-N,N′-dicyclohexyldiglycolamide (DMDCHDGA) into bis(trifluoromethylsulfoyl)imide-based ionic liquids (ILs), namely 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([C4mim][Tf2N]), 1-octyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([C8mim][Tf2N]), benzyltriethylammonium bis(trifluoromethylsulfonyl)imide ([N222Bn][Tf2N]) methyltrioctylammonium bis(trifluoromethylsulfonyl)imide ([N1888][Tf2N]), and their mixtures with molecular organic diluent 1,2-dichloroethane (DCE), is studied. DMDCHDGA has been shown to interact with components of the IL [C4mim][Tf2N]. The effect of HNO3 concentration in the aqueous phase on the extraction of Ln(III) ions is studied. The stoichiometry of the extracted complexes is determined, and the mechanism of Ln(III) extraction in a system with [C4mim][Tf2N] is discussed. It is shown that the efficiency and intragroup selectivity of the extraction of Ln(III) ions with DMDCHDGA into [C4mim][Tf2N] is significantly higher than when using its symmetric analog TODGA. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The effect of DMDCHDGA concentration in the organic phase on the distribution ratio of Tf<sub>2</sub>N<sup>−</sup> anion between 0.01 M C<sub>4</sub>mimTf<sub>2</sub>N solutions in DCE and water.</p>
Full article ">Figure 2
<p>IR spectra of DMDCHDGA (1), IL (2), and 0.1 M DMDCHDGA/IL solution (3). The most intense peak of the ligand (curve 1) is marked with the * symbol on the IR spectrum of the solution.</p>
Full article ">Figure 3
<p>(<b>a,b</b>) Fragments of IR spectra of IL (curve 2) and 0.1 M DMDCHDGA/IL solution (curve 3).</p>
Full article ">Figure 4
<p>IR spectra of solutions 0.1 M L/IL (1), 0.5 M DMDCHDGA/IL (2), and equimolar mixture of DMDCHDGA and IL (3).</p>
Full article ">Figure 5
<p>Extraction of HTf<sub>2</sub>N with 0.01 M DMDCHDGA solutions in DCE as a function of equilibrium HTf<sub>2</sub>N concentration in the aqueous phase.</p>
Full article ">Figure 6
<p>IR spectrum of the complex DMDCHDGA-HTf<sub>2</sub>N.</p>
Full article ">Figure 7
<p>The effect of DMDCHDGA concentration in the organic phase on the distribution ratio of Tf<sub>2</sub>N<sup>−</sup> anion between 0.01 M C<sub>4</sub>mimTf<sub>2</sub>N solutions in DCE and 0.1 M HNO<sub>3</sub> solutions.</p>
Full article ">Figure 8
<p>The effect of HNO<sub>3</sub> and HCl concentration in the aqueous phase on the distribution ratio of Tf<sub>2</sub>N<sup>−</sup> anion between 0.02 M C<sub>4</sub>mimTf<sub>2</sub>N solutions in DCE and 0.02 M C<sub>4</sub>mimTf<sub>2</sub>N solutions in DCE containing 0.02 M DMDCHDGA.</p>
Full article ">Figure 9
<p>The effect of HNO<sub>3</sub> concentration<sup>−</sup> in the aqueous phase on the transfer of Tf<sub>2</sub>N<sup>−</sup> ions into the aqueous phase from IL phase in the presence of DMDCHDGA.</p>
Full article ">Figure 10
<p>IR spectra of samples 3 (0.1 M DMDCHDGA/IL) and 4 (0.1 M DMDCHDGA/IL//3 M HNO<sub>3</sub>).</p>
Full article ">Figure 11
<p>The extraction of lanthanides(III) from 3 M HNO<sub>3</sub> solutions with 0.01 M DMDCHDGA and TODGA solutions in [C<sub>4</sub>mim][Tf<sub>2</sub>N] and DCE.</p>
Full article ">Figure 12
<p>The effect of HNO<sub>3</sub> concentrations in the aqueous phase on the extraction of Ln(III) with 0.01 M solutions of DMDCHDGA in [C<sub>4</sub>mim][Tf<sub>2</sub>N].</p>
Full article ">Figure 13
<p>The effect of NO<sub>3</sub><sup>−</sup> (HNO<sub>3</sub> + NH<sub>4</sub>NO<sub>3</sub>) concentrations in the aqueous phase on the extraction of Ln(III) with 0.01 M solutions of DMDCHDGA in [C<sub>4</sub>mim][Tf<sub>2</sub>N]. [H<sup>+</sup>] = 2 M.</p>
Full article ">Figure 14
<p>The effect of H<sup>+</sup> concentrations in the aqueous phase on the extraction of Ln(III) with 0.01 M solutions of DMDCHDGA in [C<sub>4</sub>mim][Tf<sub>2</sub>N]. [NO<sub>3</sub><sup>−</sup>] = 5 M.</p>
Full article ">Figure 15
<p>The effect of DMDCHDGA concentration in [C<sub>4</sub>mim][Tf<sub>2</sub>N] on the extraction of lanthanides(III) from 3 M HNO<sub>3</sub> solutions.</p>
Full article ">Figure 16
<p>IR spectra of samples 4 (0.1 DMDCHDGA/IL//3 M HNO3) and 5 (0.1 DMDCHDGA/IL//3 M HNO<sub>3</sub> + Eu<sup>3+</sup>).</p>
Full article ">Figure 17
<p>Difference IR spectra of samples 5 and 4 (curve 1) and 7 and 6 (curve 2).</p>
Full article ">Figure 18
<p>The extraction of lanthanides(III) from 3 M HNO<sub>3</sub> solutions with 0.01 M DMDCHDGA solutions in undiluted ILs.</p>
Full article ">Figure 19
<p>The effect of HNO<sub>3</sub> concentrations in the aqueous phase on the extraction of Eu(III) with 0.01 M solutions of DMDCHDGA in DCE and DCE containing 0.01 M C<sub>4</sub>mimTf<sub>2</sub>N.</p>
Full article ">Figure 20
<p>The extraction of lanthanides(III) from 3 M HNO<sub>3</sub> and 3 M HCl solutions with 0.01 M DMDCHDGA solutions in DCE and DCE containing 0.01 M C<sub>4</sub>mimTf<sub>2</sub>N. At the Ln(III) extraction from 3 M HCl solutions in the absence of C<sub>4</sub>mimTf<sub>2</sub>N in the organic phase, the <span class="html-italic">D</span><sub>Ln</sub> values are &lt;10<sup>−2</sup>.</p>
Full article ">Figure 21
<p>The extraction of lanthanides(III) from 3 M HNO<sub>3</sub> solutions with 0.01 M DMDCHDGA and TODGA solutions in DCE containing 0.01 M ionic liquids.</p>
Full article ">Scheme 1
<p>The structures of the studied DGAs and ILs.</p>
Full article ">Scheme 2
<p>The scheme of N,N′-dimethyl-N,N′-dicyclohexyl-3-oxadiglycolamide (DMDCHDGA) synthesis.</p>
Full article ">Scheme 3
<p>The scheme of benzyltriethylammonium bis(trifluoromethylsulfonyl)imide synthesis.</p>
Full article ">
17 pages, 4784 KiB  
Article
Synthesis of an Ethylenediaminetetraacetic Acid-like Ligand Based on Sucrose Scaffold and Complexation and Proton Relaxivity Studies of Its Gadolinium(III) Complex in Solution
by Ping Zhang, Cécile Barbot, Ramakrishna Gandikota, Cenxiao Li, Laura Gouriou, Géraldine Gouhier and Chang-Chun Ling
Molecules 2024, 29(19), 4688; https://doi.org/10.3390/molecules29194688 - 3 Oct 2024
Viewed by 798
Abstract
Sucrose constitutes a non-toxic, biodegradable, low-cost and readily available natural product. To expand its utility, we developed total synthesis for a ligand based on a sucrose scaffold for potential use as a metal chelation agent. The designed target (compound 2) has a [...] Read more.
Sucrose constitutes a non-toxic, biodegradable, low-cost and readily available natural product. To expand its utility, we developed total synthesis for a ligand based on a sucrose scaffold for potential use as a metal chelation agent. The designed target (compound 2) has a metal-chelating functionality at both the C-6 and C-6’ positions, which can provide a first coordination sphere of eight valencies. The designed total synthesis was highly efficient. To demonstrate the utility of the ligand, we studied its complexation with Gd(III). Using potentiometric titration and high-resolution mass spectrometry, we confirmed the formation of a 1:1 complex with Gd(III), which has a respectable formation constant of ~1013.4. Further NMR relaxivity studies show that the Gd(III) complex has a relaxivity (r1) of 7.6958 mmol−1 s−1. Full article
(This article belongs to the Special Issue Exclusive Feature Papers on Molecular Structure)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structures of sucrose disaccharide (<b>1</b>) and newly designed synthetic ligand (<b>2</b>) using sucrose as a scaffold.</p>
Full article ">Figure 2
<p>Recorded NMR spectra and signal assignments of compound <b>2</b> in D<sub>2</sub>O. Top: 1D <sup>1</sup>H spectrum (400 MHz); bottom: 1D <sup>13</sup>C spectrum with the expansion of the ~48.0–79.5 ppm region.</p>
Full article ">Figure 3
<p>Recorded 2D <sup>1</sup>H-<sup>13</sup>C heteronuclear correlation NMR spectrum (400 MHz) of compound <b>2</b> in D<sub>2</sub>O.</p>
Full article ">Figure 4
<p>Potentiometric titrations of compound <b>2</b> (<b>left</b>). L: 0.5–3 µmol; extra HCl: 0–34.5 µmol in 0.1 M NMe<sub>4</sub>Cl; total initial volume: 4.0 mL. Burette: [NMe<sub>4</sub>OH] = 0.05 M, and calculated distribution curves (<b>right</b>) of different protonated species of compound <b>2</b> at different pHs: [compound <b>2</b>] = 5.0 × 10<sup>−4</sup> M.</p>
Full article ">Figure 5
<p>Electrospray ionization (negative) high-resolution spectrum (<b>top</b>) of formed 1:1 complex [GdL]<sup>−</sup> with a molecular formula of C<sub>26</sub>H<sub>34</sub>N<sub>8</sub>O<sub>17</sub>Gd [M-H]<sup>−</sup> between ligand <b>2</b> and Gd(III), and comparison to the isotope patterns of simulated mass spectrum (<b>bottom</b>).</p>
Full article ">Figure 6
<p>ATR-FTIR spectra of ligand <b>2</b> (<b>left</b>) and ligand <b>2</b>/Gd(III) complex (<b>right</b>).</p>
Full article ">Figure 7
<p>Potentiometric titrations of ligand <b>2</b> in the presence of gadolinium(III) (<b>left</b>). Seventeen titrations were performed. Ligand <b>2</b>: 2.0 µmol; extra HCl: 13.82 µmol in NMe<sub>4</sub>Cl (0.1 M); total initial volume: 4.0 mL. Burette: [NMe<sub>4</sub>OH] = 0.05 M. The calculated distribution curved of 1:1 complex during titration [Gd]t = [Ligand <b>2</b>] = 5.0 × 10<sup>−4</sup> M (<b>right</b>).</p>
Full article ">Figure 8
<p>p<span class="html-italic">Gd</span> value of Gd(III)-ligand <b>2</b> complexes as a function of pH. p<span class="html-italic">Gd</span> = −log[Gd]<sub>free</sub>, [Gd]<sub>total</sub> = 1.0 µM, and [ligand]<sub>total</sub> = 10 µM.</p>
Full article ">Figure 9
<p>The plot of 1/T<sub>1</sub> versus the concentration of ligand <b>2</b>/Gd(III) complex gives a linear line with a slope of 7.6958 mM<sup>−1</sup>s<sup>−1</sup>.</p>
Full article ">Scheme 1
<p>Synthetic route to the designed target ligand (<b>2</b>) from sucrose (<b>1</b>).</p>
Full article ">Scheme 2
<p>The major protonation states of target ligand <b>2</b> and their designated notations.</p>
Full article ">Scheme 3
<p>The three protonation states of iminodiacetic acids.</p>
Full article ">Scheme 4
<p>Suggested complexation reaction of ligand <b>2</b> with Gd(III) ion results in the release of protons.</p>
Full article ">Scheme 5
<p>Suggested dynamic formation of aquacomplex of ligand <b>2</b>/Gd(III) complex with water (pH &lt; ~7.5).</p>
Full article ">
22 pages, 3899 KiB  
Article
Physicochemical Characterization and Antimicrobial Properties of Lanthanide Nitrates in Dilute Aqueous Solutions
by Galina Kuz’micheva, Alexander Trigub, Alexander Rogachev, Andrey Dorokhov and Elena Domoroshchina
Molecules 2024, 29(17), 4023; https://doi.org/10.3390/molecules29174023 - 25 Aug 2024
Cited by 1 | Viewed by 994
Abstract
This work presents the results of studying dilute aqueous solutions of commercial Ln(NO3)3 · xH2O salts with Ln = Ce-Lu using X-ray diffraction (XRD), IR spectroscopy, X-ray absorption spectroscopy (XAS: EXAFS/XANES), and pH measurements. As a [...] Read more.
This work presents the results of studying dilute aqueous solutions of commercial Ln(NO3)3 · xH2O salts with Ln = Ce-Lu using X-ray diffraction (XRD), IR spectroscopy, X-ray absorption spectroscopy (XAS: EXAFS/XANES), and pH measurements. As a reference point, XRD and XAS measurements for characterized Ln(NO3)3 · xH2O microcrystalline powder samples were performed. The local structure of Ln-nitrate complexes in 20 mM Ln(NO3)3 · xH2O aqueous solution was studied under total external reflection conditions and EXAFS geometry was applied to obtain high-quality EXAFS data for solutions with low concentrations of Ln3+ ions. Results obtained by EXAFS spectroscopy showed significant contraction of the first coordination sphere during the dissolution process for metal ions located in the middle of the lanthanide series. It was established that in Ln(NO3)3 · xH2O solutions with Ln = Ce, Sm, Gd, Yb (c = 134, 100, 50 and 20 mM) there are coordinated and, to a greater extent, non-coordinated nitrate groups with bidentate and predominantly monodentate bonds with Ln ions, the number of which increases upon transition from cerium to ytterbium. For the first time, the antibacterial and antifungal activity of Ln(NO3)3 · xH2O Ln = Ce, Sm, Gd, Tb, Yb solutions with different concentrations and pH was presented. Cross-relationships between the concentration of solutions and antimicrobial activity with the type of Ln = Ce, Sm, Gd, Tb, Yb were established, as well as the absence of biocidal properties of solutions with a concentration of 20 mM, except for Ln = Yb. The important role of experimental conditions in obtaining and interpreting the results was noted. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental angular dependence of Ce-fluorescence yield from 20 mM aqueous solution of Ce(NO<sub>3</sub>)<sub>3</sub> · 6H<sub>2</sub>O salts (squares). The dashed line represents the calculated angular dependence of the film, in which metal ions are distributed in a layer 10 Å thick at the air/liquid interface. The energy of the incident beam was 13.6 keV.</p>
Full article ">Figure 2
<p>Dependence of <span class="html-italic">Ln</span>-O distance (R, Å) from central atom type. Red and green curves are used to plot data for solutions and solids, respectively. The red arrow indicates a “break” in the interatomic distances at Yb.</p>
Full article ">Figure 3
<p>(<b>a</b>) Average <span class="html-italic">Ln</span>-O interatomic distances (R, Å) in the structures of <span class="html-italic">Ln</span>(NO<sub>3</sub>)<sub>3</sub> · <span class="html-italic">x</span>H<sub>2</sub>O salts, according to literature data highlighting the (<b>b</b>) Sm-Dy, (<b>c</b>) Dy-Yb regions. Red dots: interatomic distances calculated from phase analysis of the commercial samples we studied.</p>
Full article ">Figure 4
<p>Diffraction patterns of water and solutions with c = 50 mM and c = 134 mM: <span class="html-italic">Ln</span>(NO<sub>3</sub>)<sub>3</sub> · <span class="html-italic">x</span>H<sub>2</sub>O with (<b>a</b>) <span class="html-italic">Ln</span> = Ce, (<b>b</b>) Sm, (<b>c</b>) Gd, (<b>d</b>) Tb, (<b>e</b>) Yb.</p>
Full article ">Figure 5
<p>(<b>a</b>) FT-IR spectra of <span class="html-italic">Ln</span>(NO<sub>3</sub>)<sub>3</sub> · <span class="html-italic">x</span>H<sub>2</sub>O solutions (<span class="html-italic">Ln</span> = Ce, Sm, Gd, Yb); (<b>c</b>) part of the spectrum (1200–1500 cm<sup>−1</sup>) of Yb(NO<sub>3</sub>)<sub>3</sub> · xH<sub>2</sub>O solution (c = 50 mM and 100 mM); ((<b>b</b>,<b>d</b>,<b>e</b>) for c = 20 mM) spectrum sections.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) FT-IR spectra of <span class="html-italic">Ln</span>(NO<sub>3</sub>)<sub>3</sub> · <span class="html-italic">x</span>H<sub>2</sub>O solutions (<span class="html-italic">Ln</span> = Ce, Sm, Gd, Yb); (<b>c</b>) part of the spectrum (1200–1500 cm<sup>−1</sup>) of Yb(NO<sub>3</sub>)<sub>3</sub> · xH<sub>2</sub>O solution (c = 50 mM and 100 mM); ((<b>b</b>,<b>d</b>,<b>e</b>) for c = 20 mM) spectrum sections.</p>
Full article ">Figure 6
<p>Relationship between the pH value and the concentration of <span class="html-italic">Ln</span>(NO<sub>3</sub>)<sub>3</sub> · <span class="html-italic">x</span>H<sub>2</sub>O solution with (<b>a</b>) <span class="html-italic">x</span> = 6 for <span class="html-italic">Ln</span> = Ce, Sm, Gd and <span class="html-italic">x</span> is unknown for Yb; (<b>b</b>) <span class="html-italic">x</span> = 6 for <span class="html-italic">Ln</span> = Ce, Sm, Tb, Gd and <span class="html-italic">x</span> is unknown for Yb (pH measurement error is ±0.03).</p>
Full article ">Figure 7
<p>Relationship between the growth inhibition zone (D, mm) and (<b>a</b>) concentration (c, mM) and (<b>b</b>) pH of <span class="html-italic">Ln</span>(NO<sub>3</sub>)<sub>3</sub> · <span class="html-italic">x</span>H<sub>2</sub>O solutions (<span class="html-italic">Ln</span> = Ce, Sm, Gd, Tb, Yb). Convergence of results based on three independent measurements; D ± 0.02 mm.</p>
Full article ">
16 pages, 4970 KiB  
Article
Spectroscopic Studies of Lanthanide(III) Complexes with L-Malic Acid in Binary Systems
by Michał Zabiszak, Justyna Frymark, Jakub Grajewski and Renata Jastrzab
Int. J. Mol. Sci. 2024, 25(17), 9210; https://doi.org/10.3390/ijms25179210 - 25 Aug 2024
Viewed by 875
Abstract
Binary systems of lanthanide ions (La, Nd, Gd, Ho, Tb, and Lu) with L-malic acid in molar ratios of 1:1 and 1:2 were studied. This study was carried out in aqueous solutions, and the composition of the formed complexes was confirmed using computer [...] Read more.
Binary systems of lanthanide ions (La, Nd, Gd, Ho, Tb, and Lu) with L-malic acid in molar ratios of 1:1 and 1:2 were studied. This study was carried out in aqueous solutions, and the composition of the formed complexes was confirmed using computer data analysis. The overall stability constants of the complexes and the equilibrium constants of the reaction were determined. The effect of ligand concentration on the composition of the internal coordination sphere of the central atom was observed. Changes in the coordination sphere of lanthanide ions were confirmed by spectroscopic measurements. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Distribution diagram for L-malic acid.</p>
Full article ">Figure 2
<p>Distribution diagram for the equimolar systems studied: (<b>a</b>) La(III)/L-malic acid; (<b>b</b>) Nd(III)/L-malic acid; (<b>c</b>) Gd(III)/L-malic acid; (<b>d</b>) Tb(III)/L-malic acid; (<b>e</b>) Ho(III)/L-malic acid; (<b>f</b>) and Lu(III)/L-malic acid.</p>
Full article ">Figure 3
<p>Distribution diagram for the systems studied with excess of L-malic acid: (<b>a</b>) La(III)/L-malic acid; (<b>b</b>) Nd(III)/L-malic acid; (<b>c</b>) Gd(III)/L-malic acid; (<b>d</b>) Tb(III)/L-malic acid; (<b>e</b>) Ho(III)/L-malic acid; and (<b>f</b>) Lu(III)/L-malic acid.</p>
Full article ">Figure 4
<p>Possible configurations of L-malic acid with lanthanide ions in formed complexes.</p>
Full article ">Figure 5
<p>Emission spectra of the systems: (<b>a</b>) Tb(III)/L-malic acid (1:1 ratio); (<b>b</b>) Tb(III)/L-malic acid (1:2 ratio).</p>
Full article ">Figure 6
<p>IR spectra for the equimolar systems studied: (<b>a</b>) La(III)/L-malic acid; (<b>b</b>) Nd(III)/L-malic acid; (<b>c</b>) Gd(III)/L-malic acid; (<b>d</b>) Tb(III)/L-malic acid; (<b>e</b>) Ho(III)/L-malic acid; and (<b>f</b>) Lu(III)/L-malic acid.</p>
Full article ">Figure 7
<p>IR spectra for the systems studied with excess of L-malic acid: (<b>a</b>) La(III)/L-malic acid; (<b>b</b>) Nd(III)/L-malic acid; (<b>c</b>) Gd(III)/L-malic acid; (<b>d</b>) Tb(III)/L-malic acid; (<b>e</b>) Ho(III)/L-malic acid; and (<b>f</b>) Lu(III)/L-malic acid.</p>
Full article ">Figure 8
<p>The CD spectra of the 1:1 systems: (<b>a</b>) La/Mal at pH 3.0, 6.3, and 8.9; (<b>b</b>) Ho/Mal at pH 4.3, 6.2, 7.7, and 10.2.</p>
Full article ">Figure 9
<p>The CD spectra of the 1:2 systems: (<b>a</b>) La/Mal at pH 2.5, 6.3, and 9.2; (<b>b</b>) Ho/Mal at pH 4.4, 6.3, and 8.0.</p>
Full article ">
18 pages, 4858 KiB  
Article
Sandwich d/f Heterometallic Complexes [(Ln(hfac)3)2M(acac)3] (Ln = La, Pr, Sm, Dy and M = Co; Ln = La and M = Ru)
by Cristian Grechi, Silvia Carlotto, Massimo Guelfi, Simona Samaritani, Lidia Armelao and Luca Labella
Molecules 2024, 29(16), 3927; https://doi.org/10.3390/molecules29163927 - 20 Aug 2024
Cited by 1 | Viewed by 1004
Abstract
Sandwich d/f heterometallic complexes [(Ln(hfac)3)2M(acac)3] (Ln = La, Pr, Sm, Dy and M = Co; Ln = La and M = Ru) were prepared in strictly anhydrous conditions reacting the formally unsaturated fragment [Ln(hfac)3] and [...] Read more.
Sandwich d/f heterometallic complexes [(Ln(hfac)3)2M(acac)3] (Ln = La, Pr, Sm, Dy and M = Co; Ln = La and M = Ru) were prepared in strictly anhydrous conditions reacting the formally unsaturated fragment [Ln(hfac)3] and [M(acac)3] in a 2-to-1 molar ratio. These heterometallic complexes are highly sensitive to moisture. Spectroscopic observation revealed that on hydrolysis, these compounds yield dinuclear heterometallic compounds [Ln(hfac)3M(acac)3], prepared here for comparison purposes only. Quantum mechanical calculations supported, on the one hand, the hypothesis on the geometrical arrangement obtained from ATR-IR and NMR spectra and, on the other hand, helped to rationalize the spontaneous hydrolysis reaction. Full article
Show Figures

Figure 1

Figure 1
<p>ATR-IR spectra of heterodinuclear [Ln(hfac)<sub>3</sub>Co(acac)<sub>3</sub>] (<b>1Ln</b>) complexes: Ln = La, (black); Pr, (blue); Sm, (green) and Dy (orange).</p>
Full article ">Figure 2
<p>The <sup>1</sup>H (<b>left</b>) and <sup>19</sup>F (<b>right</b>) NMR of [La(hfac)<sub>3</sub>Co(acac)<sub>3</sub>], <b>1La</b>.</p>
Full article ">Figure 3
<p>ATR-IR spectra of [(Ln(hfac)<sub>3</sub>)<sub>2</sub>Co(acac)<sub>3</sub>] (<b>2Ln</b>): La, (black); Pr, (blue); Sm, (green), Dy, (red).</p>
Full article ">Figure 4
<p>ATR-IR spectra of [(La(hfac)<sub>3</sub>)<sub>2</sub>Co(acac)<sub>3</sub>] (black) and [La(hfac)<sub>3</sub>Co(acac)<sub>3</sub>] (blue).</p>
Full article ">Figure 5
<p>ATR-IR in the range of 1700–1350 cm<sup>−1</sup>: (i) [(La(hfac)<sub>3</sub>)<sub>2</sub>Co(acac)<sub>3</sub>] (black); (ii) [(La(hfac)<sub>3</sub>)<sub>2</sub>Co(acac)<sub>3</sub>] after 1 h air exposure (blue); (iii) [La(hfac)<sub>3</sub>Co(acac)<sub>3</sub>] (green). Bands increasing (green *) or decreasing (black *) with time are marked.</p>
Full article ">Figure 6
<p>The <sup>1</sup>H (<b>left</b>) and <sup>19</sup>F (<b>right</b>) NMR spectra of [(La(hfac)<sub>3</sub>)<sub>2</sub>Co(acac)<sub>3</sub>] in CDCl<sub>3</sub>.</p>
Full article ">Figure 7
<p>The <sup>1</sup>H and <sup>19</sup>F NMR spectra of [(La(hfac)<sub>3</sub>)<sub>2</sub>Co(acac)<sub>3</sub>] in CDCl<sub>3</sub>: (i) immediately after preparation (black) and (ii) after a few days (brown).</p>
Full article ">Figure 8
<p>Optimized structure of [(La(hfac)<sub>3</sub>)<sub>2</sub>Co(acac)<sub>3</sub>] (<b>2La</b>). The green, grey, cyan, red and violet spheres are F, C, La, O and Co atoms, respectively. H atoms are omitted for clarity. Level of theory: PBE def2/JK, DFT-D3.</p>
Full article ">Figure 9
<p>The stability trend, as ∆G values, for the dinuclear [Ln(hfac)<sub>3</sub>Co(acac)<sub>3</sub>] and trinuclear [(Ln(hfac)<sub>3</sub>)<sub>2</sub>Co(acac)<sub>3</sub>] complexes (Ln = La, Pr and Sm). All values are in kcal/mol.</p>
Full article ">Figure 10
<p>(<b>Left</b>): ATR-IR of [La(hfac)<sub>3</sub>Ru(acac)<sub>3</sub>] (black) and [La(hfac)<sub>3</sub>Co(acac)<sub>3</sub>] (blue). (<b>Right</b>): ATR-IR of [(La(hfac)<sub>3</sub>)<sub>2</sub>Ru(acac)<sub>3</sub>] (black) and [La(hfac)<sub>3</sub>Ru(acac)<sub>3</sub>] (blue).</p>
Full article ">Figure 11
<p>ATR-IR of (i) [(La(hfac)<sub>3</sub>)<sub>2</sub>Ru(acac)<sub>3</sub>] (black); (ii) after 4′ of air exposure (blue); (iii) after 30′ (green) and [La(hfac)<sub>3</sub>Ru(acac)<sub>3</sub>] (red). Bands increasing (red *) or decreasing (black *) with time are marked.</p>
Full article ">Figure 12
<p>Numbering of <sup>1</sup>H NMR protons in [Ln(hfac)<sub>3</sub>M(acac)<sub>3</sub>].</p>
Full article ">Figure 13
<p>Numbering of <sup>1</sup>H NMR protons in [(La(hfac)<sub>3</sub>)<sub>2</sub>Co(acac)<sub>3</sub>].</p>
Full article ">Scheme 1
<p>Heterodinuclear metal complex formation, exemplified here for a formally unsaturated [Ln(hfac)<sub>3</sub>] and [M(acac)<sub>3</sub>] with the metal M in an octahedral <span class="html-italic">tris</span>-chelate environment.</p>
Full article ">Scheme 2
<p>Heterotrinuclear f-d-f sandwich metal complex formation, exemplified here for Ln = La and M = Co or Ru in an octahedral <span class="html-italic">tris</span>-chelate environment.</p>
Full article ">
28 pages, 9422 KiB  
Review
Comprehensive Review of Synthesis, Optical Properties and Applications of Heteroarylphosphonates and Their Derivatives
by Krzysztof Owsianik, Adrian Romaniuk, Marika Turek and Piotr Bałczewski
Molecules 2024, 29(15), 3691; https://doi.org/10.3390/molecules29153691 - 4 Aug 2024
Viewed by 1412
Abstract
This review focuses on optical properties of compounds in which at least one phosphonate group is directly attached to a heteroaromatic ring. Additionally, the synthesis and other applications of these compounds are addressed in this work. The influence of the phosphonate substituent on [...] Read more.
This review focuses on optical properties of compounds in which at least one phosphonate group is directly attached to a heteroaromatic ring. Additionally, the synthesis and other applications of these compounds are addressed in this work. The influence of the phosphonate substituent on the properties of the described compounds is discussed and compared with other non-phosphorus substituents, with particular attention given to photophysical properties, such as UV-Vis absorption and emission, fluorescence quantum yield and fluorescence lifetime. Considering the presence of heteroatom, the collected material was divided into two parts, and a review of the literature of the last thirty years on heteroaryl phosphonates containing sulfur and nitrogen atoms in the aromatic ring was conducted. Full article
(This article belongs to the Special Issue Organophosphorus Chemistry: A New Perspective, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Absorbance spectra of the four Ru-oligothiophenes (Ru-<b>1</b>, brown; Ru-<b>2</b>, red; Ru-<b>3</b>, green; Ru-<b>4</b>, blue; 0.01 mM in acetonitrile) (Reprinted with permission from [<a href="#B13-molecules-29-03691" class="html-bibr">13</a>], Copyright 2007 American Chemical Society).</p>
Full article ">Figure 2
<p>Chemical structures of chalcogenorhodamine dyes <b>18</b>–<b>20</b>.</p>
Full article ">Figure 3
<p>Emission spectra of complex Cd<sub>2</sub>[OOCC<sub>5</sub>H<sub>3</sub>NPO<sub>3</sub>H]<sub>2</sub>·H<sub>2</sub>O (<b>28</b>) (<span class="html-italic">λ</span><sub>ex</sub> = 323 nm), Zn[OOCC<sub>5</sub>H<sub>4</sub>NPO<sub>3</sub>]·H<sub>2</sub>O (<b>29</b>) (λ<sub>ex</sub> = 365 nm) and free ligand <b>27</b> (λ<sub>ex</sub> = 521 nm) (reprinted with permission from [<a href="#B28-molecules-29-03691" class="html-bibr">28</a>], Copyright Taylor and Francis Ltd., <a href="http://www.tandfonline.com" target="_blank">http://www.tandfonline.com</a>).</p>
Full article ">Figure 4
<p>Chemical structures of iridium complexes <b>39</b>–<b>42</b>.</p>
Full article ">Figure 5
<p>Chemical structures of complexes <b>43</b>–<b>45</b>.</p>
Full article ">Figure 6
<p>Chemical structures of ligands <b>46</b>–<b>52</b>.</p>
Full article ">Figure 7
<p>Chemical structures of <span class="html-italic">N,O</span>-donor <span class="html-italic">N</span>-heterocyclic aromatic diphosphonate ligands <b>53</b>–<b>55</b>.</p>
Full article ">Figure 8
<p>Chemical structures of ligands <b>56</b>–<b>58</b>.</p>
Full article ">Figure 9
<p>Chemical structures of ruthenium complexes <b>59</b>–<b>61</b>.</p>
Full article ">Figure 10
<p>Chemical structures of complexes <b>65</b>–<b>68</b>.</p>
Full article ">Figure 11
<p>Chemical structures of iridium(III) complexes <b>84</b> and <b>85</b>.</p>
Full article ">Figure 12
<p>Chemical structures of platinum complexes <b>86</b> and <b>87</b>.</p>
Full article ">Figure 13
<p>Chemical structures of ligands <b>91</b>–<b>94</b> and equilibrium <b>A</b>–<b>D</b> of <b>92</b> at different pH.</p>
Full article ">Figure 14
<p>Chemical structures of complexes <b>95</b>–<b>97</b>.</p>
Full article ">Figure 15
<p>Chemical structures of the sensitizer/donor <b>106</b> and the dyad <b>107</b>/<b>108</b>.</p>
Full article ">Figure 16
<p>Proposed electron fluxes in the illuminated heterotriad <b>106</b>|TiO<sub>2</sub> on SnO<sub>2</sub> (Reprinted with permission from [<a href="#B53-molecules-29-03691" class="html-bibr">53</a>]).</p>
Full article ">Figure 17
<p>Chemical structure of complex <b>112</b>.</p>
Full article ">Figure 18
<p>Chemical structure of 2-pyrazinephosphonic acid <b>113</b>.</p>
Full article ">Figure 19
<p>Chemical structures of pyrazine-functionalized calix [<a href="#B4-molecules-29-03691" class="html-bibr">4</a>]arene ligands <b>117</b>–<b>119</b>, and a schematic presentation of metal co-ordination.</p>
Full article ">Figure 20
<p>Absorption spectra of ligands <b>117</b>–<b>119</b> in BumimTf<sub>2</sub>N (conc.: 0.25 mmol/L at 25 °C) (reprinted with permission from [<a href="#B58-molecules-29-03691" class="html-bibr">58</a>], Copyright The Royal Society of Chemistry 2015).</p>
Full article ">Figure 21
<p>Chemical structures of TBCA (<b>123</b>) and TPCA (<b>124</b>).</p>
Full article ">Figure 22
<p>Device configuration of multilayer PLEDs and molecular structures of used materials (GPF—green-emitting polyfluorene; PEDOT:PSS—poly(3,4-ethylenedioxythiophene)-poly(styrenesulfonate); ETL—electron-transporting layer; EML—emitting layer; ITO—indium tin oxide) (Used with permission of The Royal Society of Chemistry, from [<a href="#B63-molecules-29-03691" class="html-bibr">63</a>], Copyright 2016; permission conveyed through Copyright Clearance Center, Inc.).</p>
Full article ">Scheme 1
<p>Synthesis of bithiophenes and tetrathiophenes <b>1</b>–<b>4</b>.</p>
Full article ">Scheme 2
<p>Synthesis of chalcogenophosphonato-substituted bithiophenes. Reagents and conditions: (i) 2,2′-biphenol or 2,2-dimethyl-1,3-propanediol, toluene, reflux; (ii) (a) BuLi, −78 °C, THF; (b) Ph<sub>2</sub>PCl, 0 °C—rt, 2 h; (iii) (X = O) H<sub>2</sub>O<sub>2</sub>·urea complex or N(O)Me<sub>3</sub>; (X = S) S<sub>8</sub>; (X = Se) Se powder, DCM, rt.</p>
Full article ">Scheme 3
<p>Synthesis of bis(aryl- and heteroaryl)phosphonic acids <b>13</b>–<b>17</b>.</p>
Full article ">Scheme 4
<p>The synthesis of compound <b>18</b>.</p>
Full article ">Scheme 5
<p>Synthesis of the ligand <b>30</b>.</p>
Full article ">Scheme 6
<p>Synthesis of ligands <b>31</b> and <b>32</b>.</p>
Full article ">Scheme 7
<p>Synthetic protocol for the preparation of ligands <b>35</b> and <b>36</b>.</p>
Full article ">Scheme 8
<p>Synthesis of iridium complex <b>38</b> from complex <b>37</b>.</p>
Full article ">Scheme 9
<p>Synthesis of the complex <b>63</b>.</p>
Full article ">Scheme 10
<p>Synthesis of complex <b>64</b>.</p>
Full article ">Scheme 11
<p>General synthesis of ligands <b>72</b>.</p>
Full article ">Scheme 12
<p>Synthesis and spectroscopic data of complexes <b>73</b>–<b>80</b> (absorption maximum measured in 0.1 N sulfuric acid; emission maximum measured for the anionic forms of the complexes at room temperature in deaerated methanol; emission lifetimes for the anionic forms of the complexes).</p>
Full article ">Scheme 13
<p>Synthesis of complex <b>82</b>.</p>
Full article ">Scheme 14
<p>Synthesis of the dimer <b>83</b> and the regeneration of complex <b>82</b>.</p>
Full article ">Scheme 15
<p>Synthesis and chemical structures of phosphonates <b>88</b>–<b>90</b>.</p>
Full article ">Scheme 16
<p>The synthesis of terpyridyl ligand <b>99</b> for the identification of ions Fe(II), Fe(III), Ru(III) and Zn(II) in dilute aqueous solutions.</p>
Full article ">Scheme 17
<p>Synthetic strategies for the synthesis of Ru complex <b>104</b>.</p>
Full article ">Scheme 18
<p>Synthesis of complex sensitizer/donor type <b>105</b>.</p>
Full article ">Scheme 19
<p>Synthesis of complex <b>110</b>.</p>
Full article ">Scheme 20
<p>Synthesis of the phosphopyrazole <b>120</b> and the pyridyl amino amide <b>121</b>.</p>
Full article ">Scheme 21
<p>Synthesis of the aminophenoxazinone <b>122</b>.</p>
Full article ">Scheme 22
<p>Synthesis of phosphonate <b>127</b> (TPPO).</p>
Full article ">
17 pages, 5153 KiB  
Article
Development of Broad-Range Microbial Minimal Culture Medium for Lanthanide Studies
by Gianmaria Oliva, Giovanni Vigliotta, Luca Di Stasio, Ermanno Vasca and Stefano Castiglione
Microorganisms 2024, 12(8), 1531; https://doi.org/10.3390/microorganisms12081531 - 26 Jul 2024
Viewed by 1006
Abstract
Rare Earth Elements (REE), also known as Lanthanides (Ln3+), are a group of 17 elements showing peculiar physical and chemical properties. Unlike technological applications, very little is known about the physiological role and toxicity of Ln3+ on biological systems, in [...] Read more.
Rare Earth Elements (REE), also known as Lanthanides (Ln3+), are a group of 17 elements showing peculiar physical and chemical properties. Unlike technological applications, very little is known about the physiological role and toxicity of Ln3+ on biological systems, in particular on microorganisms (e.g., bacteria), which represent the most abundant domains on our planet. Up to now, very limited studies have been conducted due to Ln3+ precipitation with some anions commonly present in the culture media. Therefore, the development of a minimal medium is essential to allow the study of Ln3+-microbial interactions, limiting considerably the precipitation of insoluble salts. In this regard, a new minimal culture medium capable of solubilizing large amounts of Ln3+ and allowing the growth of different microbial taxa was successfully developed. Assays have shown that the medium is capable of solubilizing Ln3+ up to 100 times more than other common culture media and allowing the growth of 63 bacteria and 5 fungi. The kinetic growth of one yeast and one Gram-positive bacterium has been defined, proving to support superior growth and biomass compared to other commonly used minimal media. Moreover, the sensitivity and uptake/absorption of a Bacillus stratosphericus strain were tested, highlighting its capability to tolerate concentrations up to 10 mM of either Cerium, Gadolinium or Lanthanum and accumulate different quantities of the three. Full article
(This article belongs to the Section Environmental Microbiology)
Show Figures

Figure 1

Figure 1
<p>Distribution diagram of the hydrolytic species of the Ce<sup>3+</sup> ion at 1.0 mM [Ce<sup>3+</sup>].</p>
Full article ">Figure 2
<p>Distribution diagram of 1.0 mM Ce<sup>3+</sup> in the presence of 3.0 mM total phosphate and 40.0 mM total citrate up to pH 7.36, when the formation of solid CePO<sub>4</sub> is observed.</p>
Full article ">Figure 3
<p>White fluffy cloud precipitate observable in the test-tubes at 1.0 mM of Ce<sup>3+</sup> in CD, DF, and DM culture media.</p>
Full article ">Figure 4
<p>Bacterial cultivation on different agarized culture media after 48 h of incubation.</p>
Full article ">Figure 5
<p>Fungi cultivation and growth on different agarized culture media after 5 days of incubation.</p>
Full article ">Figure 6
<p>Growth curve of <span class="html-italic">B. stratosphericus</span> in MCML and Davis and Mingioli (DM) media.</p>
Full article ">Figure 7
<p>Growth curve of <span class="html-italic">D. hansenii</span> in MCML and Czapek Dox (CD) media.</p>
Full article ">Figure 8
<p>Ln<sup>3+</sup> accumulation capability of <span class="html-italic">B. stratosphericus</span>, expressed as μg · mg<sup>−1</sup> of dry weight.</p>
Full article ">
17 pages, 10333 KiB  
Article
Development of a Novel, Easy-to-Prepare, and Potentially Valuable Peptide Coupling Technology Utilizing Amide Acid as a Linker
by Yaling Wang, Fan Yang and Hongyan Li
Pharmaceuticals 2024, 17(8), 981; https://doi.org/10.3390/ph17080981 - 24 Jul 2024
Viewed by 843
Abstract
The process of synthesizing radionuclide-coupled drugs, especially shutdown technology that links bipotent chelators with biomolecules, utilizes traditional coupling reactions, including emerging click chemistry; these reactions involve different drawbacks, such as complex and cumbersome reaction steps, long reaction times, and the use of catalysts [...] Read more.
The process of synthesizing radionuclide-coupled drugs, especially shutdown technology that links bipotent chelators with biomolecules, utilizes traditional coupling reactions, including emerging click chemistry; these reactions involve different drawbacks, such as complex and cumbersome reaction steps, long reaction times, and the use of catalysts at various pH values, which can negatively impact the effects of the chelating agent. To address the above problems in this study, This research designed a novel bipotent chelator coupled with peptides. In the present study, dichloromethane was used as a solvent, and the reaction was conducted at room temperature for 12 h. A one-step ring-opening method was employed to introduce the coupling functional group of tridentate amide acid. The coupling materials consisted of the amino active site of the peptide and diethylene glycol anhydride. In this paper, this study explored the reactions between different equivalents of acid anhydride coupled to the peptide (peptide sequence: HLRKLRKR) and determined that the maximum conversion of the peptide feedstock was 87%. To determine the selectivity of the reaction sites in this polypeptide, This study identified the peptide sequence at the reaction site using nuclear magnetic resonance (NMR) and liquid chromatography–mass spectrometry (LC-MS). For the selected peptide, the first reactive site was on the terminal amino group, followed by the amino group on the tetra- and hepta-lysine side chains. The tridentate amic acid framework functions as a chelating agent, capable of binding a range of lanthanide ions. This significantly reduces and optimizes the time and cost associated with synthesizing radionuclide-coupled drugs. Full article
(This article belongs to the Section Biopharmaceuticals)
Show Figures

Figure 1

Figure 1
<p>Liquid chromatograms of the reactions of different equivalents of acid anhydride with peptides.</p>
Full article ">Figure 2
<p>Histidine site number.</p>
Full article ">Figure 3
<p>Spectrum of <sup>1</sup>H NMR spectrum of raw peptides.</p>
Full article ">Figure 4
<p>Spectrum of <sup>1</sup>H-<sup>13</sup>C HSQC NMR spectrum of raw peptides.</p>
Full article ">Figure 5
<p>Spectrum of <sup>1</sup>H-<sup>1</sup>H COSY NMR spectrum of raw peptides.</p>
Full article ">Figure 6
<p>Spectrum of <sup>1</sup>H-<sup>13</sup>C HMBC NMR spectrum of raw peptides.</p>
Full article ">Figure 7
<p>Spectrum of <sup>13</sup>C-NMR spectrum of raw peptides.</p>
Full article ">Figure 8
<p>Chemical shifts of important structures in raw polypeptides.</p>
Full article ">Figure 9
<p>Coupling functional group label.</p>
Full article ">Figure 10
<p>Spectrum of <sup>1</sup>H-NMR spectrum of peptide product coupled to an anhydride molecule.</p>
Full article ">Figure 11
<p>Spectrum of <sup>1</sup>H-<sup>13</sup>C HSQC NMR spectrum of peptide product coupled to an anhydride molecule.</p>
Full article ">Figure 12
<p>Spectrum of <sup>1</sup>H-<sup>13</sup>C HMBC NMR spectrum of peptide product coupled to an anhydride molecule.</p>
Full article ">Figure 13
<p>Spectrum of <sup>1</sup>H-<sup>1</sup>HCOSY NMR spectrum of peptide product coupled to an anhydride molecule.</p>
Full article ">Figure 14
<p>Spectrum of <sup>13</sup>C-NMR spectrum of peptide product coupled to an anhydride molecule.</p>
Full article ">Figure 15
<p>The chemical shift values of important structures in peptide products coupled with an anhydride molecule.</p>
Full article ">Figure 16
<p>The structure of the polypeptide with associated equations.</p>
Full article ">
17 pages, 4987 KiB  
Article
A Series of Lanthanide Coordination Polymers as Luminescent Sensors for Selective Detection of Inorganic Ions and Nitrobenzene
by Miao Wu, Juan Song, Yun-Long Zhou, Hui-Hui Chen, Bo-Feng Duan, Ling-Xia Jin, Chuan-Qing Ren and Jiu-Fu Lu
Molecules 2024, 29(14), 3438; https://doi.org/10.3390/molecules29143438 - 22 Jul 2024
Cited by 1 | Viewed by 958
Abstract
Seven new lanthanide coordination polymers, namely [Ln(cpt)3H2O)]n(Ln = La (1), Pr (2), Sm (3), Eu (4), Gd (5), Dy (6), and Er (7)), which were [...] Read more.
Seven new lanthanide coordination polymers, namely [Ln(cpt)3H2O)]n(Ln = La (1), Pr (2), Sm (3), Eu (4), Gd (5), Dy (6), and Er (7)), which were synthesized under hydrothermal conditions using 4′-(4-(4-carboxyphenyloxy)phenyl)-4,2′:6′,4′-tripyridine (Hcpt) as the ligand. The crystal structures of these seven complexes were determined using single-crystal X-ray diffraction, and they were found to be isostructural, crystallizing in the triclinic P1- space group. The Ln(III) ions were nine-coordinated with tricapped trigonal prism coordination geometry. The Ln(III) cations were coordinated by carboxylic and pyridine groups from (cpt) ligands, forming one-dimensional ring-chain structures. Furthermore, the luminescent properties of complexes 17 were investigated using fluorescent spectra in the solid state. The fluorescence sensing experiments demonstrated that complex 4 exhibits high selectivity and sensitivity for detecting Co2+, Cu2+ ions, and nitrobenzene. Moreover, complex 3 shows good capability for detecting Cu2+ ions and nitrobenzene. Additionally, the sensing mechanism was also thoroughly examined through theoretical calculations. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Coordination environment of the Eu<sup>III</sup> cations in complex <b>4</b>. (<b>b</b>) The one-dimensional ring-chain structure of complex <b>4</b>. (<b>c</b>) Two-dimensional supramolecular structure of complex <b>4</b>.</p>
Full article ">Figure 1 Cont.
<p>(<b>a</b>) Coordination environment of the Eu<sup>III</sup> cations in complex <b>4</b>. (<b>b</b>) The one-dimensional ring-chain structure of complex <b>4</b>. (<b>c</b>) Two-dimensional supramolecular structure of complex <b>4</b>.</p>
Full article ">Figure 2
<p>(<b>a</b>) Emission spectra of complex <b>4</b>. (<b>b</b>) Emission spectra of complex <b>3</b>. (<b>c</b>) Emission spectra of complex <b>6</b>.</p>
Full article ">Figure 3
<p>(<b>a</b>) Luminescence decay curves of complex <b>4</b>. (<b>b</b>) Luminescence decay curves of complex <b>3</b>.</p>
Full article ">Figure 4
<p>CIE chromaticity diagrams for complex <b>4</b> and <b>3</b>.</p>
Full article ">Figure 5
<p>(<b>a</b>) Emission spectra of complex <b>4</b> in diverse metal ions. (<b>b</b>) Comparisons for intensities of complex <b>4</b> in different metal ions. (<b>c</b>,<b>e</b>) Luminescence quenching of complex <b>4</b> dispersed in DMF with gradual addition of Cu<sup>2+</sup> and Co<sup>2+</sup> ions, respectively. (<b>d</b>,<b>f</b>) The fitting of the fluorescence change value of complex <b>4</b> in Cu<sup>2+</sup> and Co<sup>2+</sup> ions, respectively.</p>
Full article ">Figure 6
<p>(<b>a</b>) Emission spectra of complex <b>3</b> in diverse metal ions. (<b>b</b>) Comparisons for intensities of complex <b>3</b> in different metal ions. (<b>c</b>) Luminescence quenching of complex <b>3</b> dispersed in DMF with gradual addition of Cu<sup>2+</sup> ion. (<b>d</b>) The fitting of the fluorescence change value of complex <b>3</b> in Cu<sup>2+</sup> ion.</p>
Full article ">Figure 7
<p>(<b>a</b>) Emission spectra of complex <b>4</b> in diverse analytes. (<b>b</b>) Comparisons for intensities of <b>4</b> in diverse analytes. (<b>c</b>) Emission spectra of complex <b>4</b> in CH<sub>3</sub>CN with different concentrations of NB.</p>
Full article ">Figure 8
<p>(<b>a</b>) Emission spectra of complex <b>3</b> in diverse analytes. (<b>b</b>) Comparisons for intensities of <b>3</b> in diverse analytes. (<b>c</b>) Emission spectra of complex <b>3</b> in DMF with different concentrations of NB.</p>
Full article ">Figure 9
<p>HOMO and LUMO energy levels of ligand and the selected analytes.</p>
Full article ">Figure 10
<p>Hirshfeld surface mapped with dnorm (<b>left</b>), shape index (<b>middle</b>), and curvedness (<b>right</b>) for complex <b>4</b>.</p>
Full article ">Figure 11
<p>Two-dimensional (2D) fingerprint plots of Hirshfeld surfaces for complex <b>4</b>.</p>
Full article ">Scheme 1
<p>Coordination modes of the (cpt)<sup>−</sup> ligand. (<b>a</b>) a bidentate coordination fashion (<b>b</b>) a tridentate coordination fashion.</p>
Full article ">
19 pages, 5775 KiB  
Article
Effect of Lanthanide Ions and Triazole Ligands on the Molecular Properties, Spectroscopy and Pharmacological Activity
by Mauricio Alcolea Palafox, Nataliya P. Belskaya, Lozan T. Todorov, Nadya G. Hristova-Avakoumova and Irena P. Kostova
Int. J. Mol. Sci. 2024, 25(14), 7964; https://doi.org/10.3390/ijms25147964 - 21 Jul 2024
Viewed by 751
Abstract
The effect of La, Ce, Pr and Nd ions on four Ln(ligand)3 complexes and at three DFT levels of calculation was analyzed. Four ligands were chosen, three of which were based on the 1,2,3-triazole ring. The DFT methods used were B3LYP, CAM-B3LYP [...] Read more.
The effect of La, Ce, Pr and Nd ions on four Ln(ligand)3 complexes and at three DFT levels of calculation was analyzed. Four ligands were chosen, three of which were based on the 1,2,3-triazole ring. The DFT methods used were B3LYP, CAM-B3LYP and M06-2X. The relationships established were between the geometric parameters, atomic charges, HOMO-LUMO energies and other molecular properties. These comparisons and trends will facilitate the synthesis of new complexes by selecting the ligand and lanthanide ion best suited to the desired property of the complex. The experimental IR and Raman spectra of Ln(2b′)3 complexes where Ln = La, Ce, Pr, Nd, Sm, Gd, Dy, Ho and Er ions have been recorded and compared to know the effect of the lanthanide ion on the complex. The hydration in these complexes was also analyzed. Additionally, the effect of the type of coordination center on the ability of an Ln(ligand)3 complex to participate in electron exchange and hydrogen transfer was investigated using two in vitro model systems—DPPH and ABTS. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the four Ln(III) complexes where R corresponds to methyl (A-complex); 1,2,3-triazole-4- (B-complex); 2-(4-chlorophenyl)-2<span class="html-italic">H</span>-1,2,3-triazole-4- (C-complex); and 2-(4-chlorophenyl)-5-(pyrrolidin-1-yl)-2<span class="html-italic">H</span>-1,2,3-triazole-4- (2b′ ligand) (Ln(2b′)<sub>3</sub> complex).</p>
Full article ">Figure 2
<p>Two conformations found by rotation around the carboxylate group. The italics numbers in green color corresponds to the three 2b′ ligands of the complex.</p>
Full article ">Figure 3
<p>Optimized structure of the La(2b′)<sub>3</sub> complex by the CAM-B3LYP method and the Lanl2dz and Cep-4g basis set.</p>
Full article ">Figure 4
<p>Labeling of the atoms in the 2b′ ligand of the complex.</p>
Full article ">Figure 5
<p>Relationships are established with the bond lengths and angles: (<b>a</b>,<b>b</b>) Between the Ln-O12 and C11-O12 bond lengths by CAM-B3LYP and M06-2X methods. (<b>c</b>) Between the Lanthanide charge and Ln-O12 bond length. (<b>d</b>) Between lanthanide charge and C11-O12 bond length. (<b>e</b>) Between lanthanide charge and C9-C11 bond length. (<b>f</b>) Between the O12-Ln-O13 angle and dipole moment of the complex.</p>
Full article ">Figure 6
<p>Relationships established with the atomic charges: (<b>a</b>,<b>b</b>) Between lanthanide charge and charge on O12 by CAM-B3LYP and M06-2X. (<b>c</b>) Between the lanthanide charge and charge on the C11 atom. (<b>d</b>) Between lanthanide charge and C9 charge. (<b>e</b>) Between the lanthanide charge and dipole moment of the complex. (<b>f</b>) Between the atomic charge on O12 and the dipole moment of the complex.</p>
Full article ">Figure 7
<p>Relationships established at the M06-2X level with the molecular properties: (<b>a</b>) Between lanthanide charge and HOMO energy. (<b>b</b>) Between the charge on O12 and LUMO energy. (<b>c</b>,<b>d</b>) Between the LUMO-HOMO energy and the lanthanide/O12 charges, respectively. (<b>e</b>,<b>f</b>) Between lanthanide charge and capacity at constant volume in all complexes and in Ln(2b′)<sub>3</sub> complex, respectively.</p>
Full article ">Figure 8
<p>Comparison of the experimental IR spectra of the Ln(2b′)<sub>3</sub> complexes with Ln = La, Ce, Pr, Nd, Sm, Gd, Dy, Ho and Er in the 3750–400 cm<sup>−1</sup> range.</p>
Full article ">Figure 9
<p>Comparison of the experimental Raman spectra of the Ln(2b′)<sub>3</sub> complexes with Ln = La, Ce, Pr, Nd, Sm, Gd, Dy, Ho and Er in the 2000–50 cm<sup>−1</sup> range.</p>
Full article ">Figure 10
<p>Impact of the ligand 2b′ and the various Ln(2b′)<sub>3</sub> complexes on (<b>a</b>) the DPPH and (<b>b</b>) ABTS in vitro model systems. Data = mean ± stdev, N = 4.</p>
Full article ">
16 pages, 1455 KiB  
Article
A Recyclable Inorganic Lanthanide Cluster Catalyst for Chemoselective Aerobic Oxidation of Thiols
by Lijun Wang, Zixuan Qin, Lingxia Chen, Xinshu Qin, Jiaman Hou, Chao Wang, Xuan Li, Hongxia Duan, Bing Fang, Minlong Wang and Jie An
Molecules 2024, 29(14), 3361; https://doi.org/10.3390/molecules29143361 - 17 Jul 2024
Cited by 1 | Viewed by 851
Abstract
Optimizing lanthanide catalyst performance with organic ligands often encounters significant challenges, including susceptibility to water or oxygen and complex synthesis pathways. To address these issues, our research focuses on developing inorganic lanthanide clusters with enhanced stability and functionality. In this study, we introduce [...] Read more.
Optimizing lanthanide catalyst performance with organic ligands often encounters significant challenges, including susceptibility to water or oxygen and complex synthesis pathways. To address these issues, our research focuses on developing inorganic lanthanide clusters with enhanced stability and functionality. In this study, we introduce the [Sm6O(OH)8(H2O)24]I8(H2O)8 cluster (Sm-OC) as a sustainable and efficient catalyst for the aerobic oxidation of thiols under heating conditions. The Sm-OC catalyst demonstrated remarkable stability, outstanding recyclability, and excellent chemoselectivity across a diverse range of functional groups in 38 different tests. Notably, it enables efficient unsymmetrical disulfide synthesis and prevents the formation of over-oxidized by-products, highlighting its superior performance. This Sm-OC catalyst provides a practical and robust tool for the precise construction of versatile disulfides, thus establishing a template for the broader use of lanthanide clusters in organic synthesis. Full article
Show Figures

Figure 1

Figure 1
<p>The structure of Sm-OC.</p>
Full article ">Figure 2
<p>Substrate scope for the aerobic oxidation of thiols. <span class="html-italic"><sup>a</sup></span> Conditions: Sm-OC (1.15 mol%) were added to a solution of thiols (0.300 mmol, 1.00 equiv.) in AcOEt (8.00 mL) under O<sub>2</sub> balloon at 70 °C and stirred for a duration of 16 h. <span class="html-italic"><sup>b</sup></span> Sm-OC (2.30 mol%) were added to a solution of thiols (0.300 mmol, 1.00 equiv.) in AcOEt (8.00 mL) under O<sub>2</sub> balloon at 70 °C and stirred for a duration of 16 h. <span class="html-italic"><sup>c</sup></span> Sm-OC (2.30 mol%) were added to a solution of thiols (0.300 mmol, 1.00 equiv.) and (3<span class="html-italic">s</span>,5<span class="html-italic">s</span>,7<span class="html-italic">s</span>)-adamantane-1-thiol (0.900 mmol, 3.00 equiv.) in AcOEt (8.00 mL) under O<sub>2</sub> balloon at 70 °C and stirred for a duration of 16 h.</p>
Full article ">
Back to TopTop