[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

Journals

Article Types

Countries / Regions

Search Results (33)

Search Parameters:
Keywords = chloride adducts

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
8 pages, 2827 KiB  
Communication
Coordination of O-Propyl-N-phenylthiocarbamate to HgI2 and the Crystallographic Characterization of an Anilinium Chloride Thiocarbamate Adduct
by Wafa Arar, Nuri Ekici, Michael Knorr, Isabelle Jourdain, Carsten Strohmann and Jan-Lukas Kirchhoff
Molbank 2024, 2024(4), M1923; https://doi.org/10.3390/M1923 - 22 Nov 2024
Viewed by 523
Abstract
In order to investigate the coordination chemistry of O-alkyl N-aryl thiocarbamate ligands, HgI2 was reacted with one equivalent of PrOC(=S)N(H)Ph L in toluene solution to afford the 1D polymeric title compound [{IHg(μ-I)}{κ1-PrOC(=S)N(H)Ph}]n CP1. The formation [...] Read more.
In order to investigate the coordination chemistry of O-alkyl N-aryl thiocarbamate ligands, HgI2 was reacted with one equivalent of PrOC(=S)N(H)Ph L in toluene solution to afford the 1D polymeric title compound [{IHg(μ-I)}{κ1-PrOC(=S)N(H)Ph}]n CP1. The formation of this iodide-bridged coordination polymer was ascertained by a single-crystal X-ray diffraction study performed at 100 K, as well as the formation of an adduct between anilinium chloride and L forming a supramolecular ribbon of composition [L(PhNH3)(Cl)]. The occurrence of anilinium chloride is due to the partial hydrolysis of L in the presence of HCl. Full article
(This article belongs to the Section Organic Synthesis and Biosynthesis)
Show Figures

Figure 1

Figure 1
<p>View of a segment of the supramolecular ribbon of PrOC(=S)NHPh•PhNH<sub>3</sub>Cl running along the <span class="html-italic">b</span> axis. Selected bond lengths (Å) and angles (deg). S–C1 1.6697(16), O–C1 1.329(2), O–C8 1.457(2), N1–C1 1.348(2), N1–C2 1.418(2), N2–C11 1.4674(19), C6–C5 1.386(3); C1–O–C8 119.18(12), O–C8–C9 112.09(14), O–C1–S 125.73(12), O–C1–N1 112.85(14), N1–C1–S 121.41(13), C1–N1–C2 131.79(15), C3–C2–N1 125.38(15). N2–H2A···Cl<sup>1</sup> 159.3, N2–H2B···Cl 171.5; N2–H2C···Cl<sup>2</sup> 162.6, N1–H1···Cl<sup>1</sup> 168.3. Symmetry transformation is used to generate equivalent atoms: <sup>1</sup>1/2-<span class="html-italic">x</span>, −1/2+<span class="html-italic">y</span>, 1/2-<span class="html-italic">z</span>; <sup>2</sup>1/2-<span class="html-italic">x</span>, 1/2 + <span class="html-italic">y</span>, 1/2-<span class="html-italic">z</span>.</p>
Full article ">Figure 2
<p><sup>1</sup>H NMR spectra (400 MHz, CDCl<sub>3</sub>) of <b>L</b> at 298 and 323 K. The * denotes CHCl<sub>3</sub>.</p>
Full article ">Figure 3
<p><sup>13</sup>C{<sup>1</sup>H} NMR spectrum (100 MHz, CDCl<sub>3</sub>) of <b>L</b> at 298 K. The * denotes CDCl<sub>3</sub>.</p>
Full article ">Figure 4
<p>View of a segment of <b>CP1</b> running along the <span class="html-italic">c</span> axis. Selected bond lengths (Å) and angles (deg). Hg–I1 2.6829(3), Hg–I2 2.9073(3), Hg–I2<sup>1</sup> 2.8421(3), Hg–S 2.5438(9), S–C5 1.710(4), O–C5 1.307(4), O–C6 1.473(4), N–C5 1.333(4), N–C9 1.430(4), C3–C7 1.390(5); I2<sup>1</sup>–Hg1–I2 103.989(9), I1– Hg1– I2 106.923(9), I1–Hg1–2<sup>1</sup> 116.402(9), S1–Hg1–I2<sup>1</sup> 98.49(2), S1–Hg1–I2 102.18(2), S1–Hg1– I1 126.21(2), Hg1<sup>2–</sup> I2-Hg 94.509(9), C5–O1–C6 120.0(3), C5–N1–C9 131.7(3), O–C5–S 121.5(3), O–C5–N 115.1(3), N–C5–S 123.4(3), O–C6–C10 105.9(3). Symmetry transformation is used to generate equivalent atoms: <sup>1</sup>+<span class="html-italic">x</span>, 1/2-<span class="html-italic">y</span>, 1/2 <span class="html-italic">z</span>; <sup>2</sup>+<span class="html-italic">x</span>, 1/2-<span class="html-italic">y</span>, −1/2+<span class="html-italic">z</span>.</p>
Full article ">Figure 5
<p>OLEX-generated view of the unit cell of <b>CP1</b>, indicating several intramolecular interactions by dotted lines (O2<b>···</b>H3C3 2.819(4) Å).</p>
Full article ">Scheme 1
<p>Examples of some <span class="html-italic">O</span>-alkyl <span class="html-italic">N</span>-aryl thiocarbamate complexes.</p>
Full article ">Scheme 2
<p>Synthesis of <b>L</b> and <b>CP1</b>.</p>
Full article ">
13 pages, 3364 KiB  
Article
Synthesis of Amorphous Cellulose Derivatives via Michael Addition to Hydroxyalkyl Acrylates for Thermoplastic Film Applications
by Hiroyuki Nagaishi, Masayasu Totani and Jun-ichi Kadokawa
Polymers 2024, 16(22), 3142; https://doi.org/10.3390/polym16223142 - 11 Nov 2024
Viewed by 864
Abstract
The aim of this study is to prepare new cellulose derivatives that show good feasibility and processability. Accordingly, in this study, we demonstrate Michael addition to hydroxyalkyl acrylates, that is, 2-hydroxyethyl and 4-hydroxybutyl acrylates (HEA and HBA, respectively), to synthesize amorphous cellulose derivatives [...] Read more.
The aim of this study is to prepare new cellulose derivatives that show good feasibility and processability. Accordingly, in this study, we demonstrate Michael addition to hydroxyalkyl acrylates, that is, 2-hydroxyethyl and 4-hydroxybutyl acrylates (HEA and HBA, respectively), to synthesize amorphous cellulose derivatives under alkaline conditions. The reactions were carried out in the presence of LiOH in ionic liquid (1-butyl-2,3-dimethylimidazolium chloride)/N,N-dimethylformamide (DMF) solvents at room temperature or 50 °C for 1 h. The Fourier transform infrared and 1H nuclear magnetic resonance (NMR) measurements of the products supported the progress of Michael addition; however, the degrees of substitution (DS) were not high (0.3–0.6 for HEA and 0.6 for HBA). The powder X-ray diffraction analysis of the products indicated their amorphous nature. The cellulosic Michael adduct from HEA with DS = 0.6 was swollen with high polar organic liquids, such as DMF. In addition to swelling with these liquids, the cellulosic Michael adduct from HBA was soluble in dimethyl sulfoxide (DMSO), leading to its 1H NMR analysis in DMSO-d6. This adduct was found to form a cast film with flexible properties from its DMSO solutions. Furthermore, films containing an ionic liquid, 1-butyl-3-methylimidazolium chloride, showed thermoplasticity. The Michael addition approach to hydroxyalkyl acrylates is quite effective to totally reduce crystallinity, leading to good feasibility and processability in cellulosic materials, even with low DS. In addition, the present thermoplastic films will be applied in practical, bio-based, and eco-friendly fields. Full article
(This article belongs to the Special Issue Polysaccharides: Synthesis, Properties and Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Preparation of cellulose solution in 1-butyl-2,3-dimethylimidazoium chloride (BDMIMCl)/<span class="html-italic">N</span>,<span class="html-italic">N</span>-dimethylformamide (DMF), (<b>b</b>) Michael addition to 2-hydroxyethyl or 4-hydroxybutyl acrylate (HEA/HBA) in the presence of LiOH in the solution, and (<b>c</b>) acetylation of produced Michael adducts in 1-butyl-3-methylimidazoium chloride (BMIMCl).</p>
Full article ">Figure 2
<p>FTIR spectra of (<b>a</b>) cellulose, (<b>b</b>) cellulosic Michael adduct from HEA (run 1), (<b>c</b>) cellulosic Michael adduct from HBA (run 3), and (<b>d</b>) dried material from alkaline hydrolysate (run 3).</p>
Full article ">Figure 3
<p><sup>1</sup>H NMR spectrum of hydrolysate of cellulosic Michael adduct from HBA (run 3) in NaOD/D<sub>2</sub>O.</p>
Full article ">Figure 4
<p>XRD profiles of (<b>a</b>) cellulose, (<b>b</b>) cellulosic Michael adduct from HEA (run 1), and (<b>c</b>) cellulosic Michael adduct from HBA (run 3).</p>
Full article ">Figure 5
<p>Photographs of mixtures of Michael adducts with DMSO and DMF after shaking at room temperature.</p>
Full article ">Figure 6
<p><sup>1</sup>H NMR spectrum of cellulosic Michael adduct from HBA (run 3) in DMSO-<span class="html-italic">d</span><sub>6</sub>.</p>
Full article ">Figure 7
<p>(<b>a</b>) <sup>1</sup>H NMR spectrum of acetylated derivative, prepared from cellulosic Michael adduct of run 3 in CDCl<sub>3</sub> and (<b>b</b>) expanded region for acetyl methyl signals.</p>
Full article ">Figure 8
<p>Preparation of cast films from solutions of cellulosic Michael adduct of run 3 containing 0 and 20 wt.% BMIMCl in DMSO and their bending performance.</p>
Full article ">Figure 9
<p>Stress-strain curves of cast films from solutions of cellulosic Michael adduct of run 3 containing 0 and 20 wt.% BMIMCl ((<b>a</b>) and (<b>b</b>), respectively) under tensile mode.</p>
Full article ">Figure 10
<p>DSC profiles of cast films from solutions of cellulosic Michael adduct of run 3, containing (<b>a</b>) 0, (<b>b</b>) 5, (<b>c</b>) 10, and (<b>d</b>) 20 wt.% BMIMCl.</p>
Full article ">Figure 11
<p>Melt-pressing experiment of cast films from solutions of cellulosic Michael adduct of run 3, containing 0–20 wt.% BMIMCl for evaluation of thermoplasticity and bending performance after melt pressing.</p>
Full article ">
14 pages, 3163 KiB  
Article
N-Oxide Coordination to Mn(III) Chloride
by Ananya Saju, Matthew R. Crawley, Samantha N. MacMillan, Pierre Le Magueres, Mark Del Campo and David C. Lacy
Molecules 2024, 29(19), 4670; https://doi.org/10.3390/molecules29194670 - 1 Oct 2024
Viewed by 1039
Abstract
We report on the synthesis and characterization of Mn(III) chloride (MnIIICl3) complexes coordinated with N-oxide ylide ligands, namely trimethyl-N-oxide (Me3NO) and pyridine-N-oxide (PyNO). The compounds are reactive and, while isolable in the [...] Read more.
We report on the synthesis and characterization of Mn(III) chloride (MnIIICl3) complexes coordinated with N-oxide ylide ligands, namely trimethyl-N-oxide (Me3NO) and pyridine-N-oxide (PyNO). The compounds are reactive and, while isolable in the solid-state at room temperature, readily decompose into Mn(II). For example, “[MnIIICl3(ONMe3)n]” decomposes into the 2D polymeric network compound complex salt [MnII(µ-Cl)3MnII(µ-ONMe3)]n[MnII(µ-Cl)3]n·(Me3NO·HCl)3n (4). The reaction of MnIIICl3 with PyNO forms varied Mn(III) compounds with PyNO coordination and these react with hexamethylbenzene (HMB) to form the chlorinated organic product 1-cloromethyl-2,3,4,5,6-pentamethylbenzene (8). In contrast to N-oxide coordination to Mn(III), the reaction between [MnIIICl3(OPPh3)2] and 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) resulted in electron transfer-forming d5 manganate of the [TEMPO] cation instead of TEMPO–Mn(III) adducts. The reactivity affected by N-oxide coordination is discussed through comparisons with other L–MnIIICl3 complexes within the context of reduction potential. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>MicroED structure of [Mn<sup>II</sup>(µ-Cl)<sub>3</sub>Mn<sup>II</sup>(µ-ONMe<sub>3</sub>)]<sub>n</sub>[Mn<sup>II</sup>(µ-Cl)<sub>3</sub>]<sub>n</sub>·(Me<sub>3</sub>NO·HCl)<sub>3n</sub> (<b>4</b>). The grey box contains the unit cell and is viewed down the b-axis (<b>left</b>) and c-axis (<b>right</b>). Selected bond lengths (Å) for the cationic chain: Mn1–Cl1 = 2.56(3); Mn1–O1 = 2.19(3); N1–O1 = 1.397(19); Mn1(µ-Cl)···Mn1(µ-O)···Mn1 = 3.27, 3.11. Selected bond lengths (Å) for the anionic chain: Mn2–Cl2 = 2.54(3); Mn2···Mn2 = 3.19. Color scheme: cyan polyhedra = [Mn<sup>II</sup>Cl<sub>3</sub>]<sup>−</sup>; green polyhedra = [Mn<sup>II</sup><sub>2</sub>Cl<sub>3</sub>(ONMe<sub>3</sub>)]<sup>+</sup>; green sphere = Cl; magenta spheres = Mn; blue spheres = N; red spheres = O; grey spheres = C.</p>
Full article ">Figure 2
<p>(<b>Left</b>) FTIR spectra of <b>3a</b> (red) and <b>4</b> (black). (<b>Right</b>) Molecular structure of <b>3b</b> with the outer sphere Cl counter anion shown (hydrogen atoms and MeCN are omitted for clarity). Selected bond lengths (Å) and angles (deg.) for <b>3a</b>: Mn1–Cl1 = 2.3225(6); Mn1–O1 = 1.901(2); Mn1–O2 = 1.896(2); Mn1–O3 = 1.991(3); Cl1–Mn1–Cl1 = 145.87(4); O1–Mn1–O2 = 177.33(11).</p>
Full article ">Figure 3
<p>Molecular structures of Mn(III) centers in (<b>a</b>) <b>5</b>, (<b>b</b>) <b>6</b>·<b>7</b>, and (<b>c</b>) <b>7</b> (full crystal structures are presented in the SI). Selected bond lengths (Å) and angles (deg.) for <b>5</b>: Mn1–Cl1 = 2.5535(4); Mn1–O1 = 1.9389(12); Mn1–O2 = 1.9293(11); Mn1–O3 = 1.9301(12); Mn1–O4 = 1.9255(11); Mn1–O5 = 2.2340(13); O1–Mn1–Cl1 = 93.54(4); O1–Mn1–O4 = 92.03(5); O4–Mn1–Cl1 = 89.60(4). Selected bond lengths (Å) and angles (deg.) for <b>6</b>·<b>7</b>: Mn1–Cl1 = 2.2804(6); Mn1–Cl2 = 2.3687(6); Mn1–Cl3 = 2.2790(6); Mn1–O1 = 1.9218(15); Mn1–O2 = 1.9194(16); Mn2–Cl4 = 2.5172(6); Mn2–Cl5 = 2.5390(6); Mn2–O3 = 1.9397(15); Mn2–O4 = 1.9309(15); Mn2–O5 = 1.9493(15); Mn2–O6 = 1.9465(15); Cl1–Mn1–Cl2 = 116.23(2); Cl2–Mn1–Cl3 = 105.08(2); Cl1–Mn1–Cl3 = 138.69(3); O1–Mn1–Cl1 = 90.58(5); O2–Mn1–Cl1 = 84.49(5); O3–Mn2–Cl4 = 91.04(5); O3–Mn2–O4 = 89.02(6); O4–Mn2–Cl4 = 91.28(5). Selected bond lengths (Å) and angles (deg.) for <b>7</b>: Mn1–Cl1 = 2.3172(8); Mn1–Cl2 = 2.3091(8); Mn1–Cl3 = 2.2961(8), Mn1–O1 = 1.912(2); Mn1–O2 = 1.916(2); Cl1–Mn1–Cl2 = 109.02(3); O1–Mn1–O2 = 168.97(9).</p>
Full article ">Figure 4
<p>Molecular structure (ellipsoids 50%) of <b>9</b> determined with XRD (H atoms and one part of disorder omitted, only one of the three identical subunits in the unit cell shown for clarity) Selected bond lengths (Å) and angles (deg.) for <b>9</b>: Mn1–Cl1 = 2.3737(5); Mn1–Cl2A = 2.3549(18); Mn1–Cl3A = 2.3728(10); Mn1–Cl4A = 2.4025(15); N1–O1 = 1.1922(17); N6–O6 = 1.191(2); Cl1–Mn1–Cl2A = 108.62(6); Cl3A–Mn1–Cl4A = 103.78(6).</p>
Full article ">Scheme 1
<p>Synthesis of Mn(III) chloride complexes.</p>
Full article ">Scheme 2
<p>Synthesis of <b>3a</b>, <b>3b</b>, and <b>4</b>.</p>
Full article ">Scheme 3
<p>Synthesis of <b>5</b>, <b>6</b>, and <b>7</b>. Small changes in reaction conditions give different products.</p>
Full article ">Scheme 4
<p>Conditions for C–H chlorination reactivity of HMB with Mn(III) chloride compounds.</p>
Full article ">Scheme 5
<p>Reaction of <b>1</b> with TEMPO.</p>
Full article ">
17 pages, 1700 KiB  
Article
The Effect of Ca, Sr, and Ba Chloride Complexes with Dibenzo-18-Crown-6 Ether as Catalysts on the Process Criteria for the Efficiency of Cumene Oxidation (the First Stage in the Chain of Polymer Composite Production)
by Nikolai V. Ulitin, Nikolay A. Novikov, Yana L. Lyulinskaya, Daria A. Shiyan, Konstantin A. Tereshchenko, Natalia M. Nurullina, Marina N. Denisova, Yaroslav O. Mezhuev and Kharlampii E. Kharlampidi
J. Compos. Sci. 2023, 7(2), 60; https://doi.org/10.3390/jcs7020060 - 6 Feb 2023
Viewed by 1750
Abstract
A study was made on the effect of Ca, Sr, and Ba chloride complexes with dibenzo-18-crown-6 ether as catalysts on the process criteria of the efficiency of industrial cumene oxidation using kinetic modeling. It is the first stage in the process chain of [...] Read more.
A study was made on the effect of Ca, Sr, and Ba chloride complexes with dibenzo-18-crown-6 ether as catalysts on the process criteria of the efficiency of industrial cumene oxidation using kinetic modeling. It is the first stage in the process chain of polymer composite production. The kinetic scheme of the process is made of classical reactions of the radical chain mechanism (reactions of initiation, chain propagation, and chain termination), molecular reactions, reactions of formation of intermediate adducts “component of the reaction mixture—catalyst” and their decomposition, as well as reactions that take into account the specifics of the catalyst used: (1) formation of planar catalyst complexes with various substances; (2) formation of acetophenone along the catalytic path; (3) hydration of the intermediate adduct “α-methylstyrene—catalyst” to the required alcohol. It is shown that the kinetic model fully reproduces the experimental time dependencies of the cumene hydroperoxide concentration in the cumene oxidation and cumene hydroperoxide decomposition. Using the kinetic model, computational experiments were carried out, as a result of which the following conclusions were made: (1) among the considered catalysts, the complex of Sr chloride with dibenzo-18-crown-6 ether should be recognized as the best, provided that it is used at temperatures of 393–413 K and an initial concentration < 2 mmol/L; (2) to ensure selectivity comparable to the selectivity of a non-catalytic process, it is necessary to conduct the catalytic process at a lowest possible initial concentration of any of the considered catalysts. Full article
(This article belongs to the Special Issue Progress in Polymer Composites, Volume II)
Show Figures

Figure 1

Figure 1
<p>The dependencies of the component concentrations of the reaction mixture in cumene oxidation versus time: dots are the experimental data, and lines are the calculation by the model; (<b>a</b>) ROOH; (<b>b</b>) ROH; (<b>c</b>) C<sub>6</sub>H<sub>5</sub>C(CH<sub>3</sub>) = CH<sub>2</sub>; (<b>d</b>) C<sub>6</sub>H<sub>5</sub>C(O)CH<sub>3</sub>; (<b>e</b>) C<sub>6</sub>H<sub>5</sub>OH.</p>
Full article ">Figure 1 Cont.
<p>The dependencies of the component concentrations of the reaction mixture in cumene oxidation versus time: dots are the experimental data, and lines are the calculation by the model; (<b>a</b>) ROOH; (<b>b</b>) ROH; (<b>c</b>) C<sub>6</sub>H<sub>5</sub>C(CH<sub>3</sub>) = CH<sub>2</sub>; (<b>d</b>) C<sub>6</sub>H<sub>5</sub>C(O)CH<sub>3</sub>; (<b>e</b>) C<sub>6</sub>H<sub>5</sub>OH.</p>
Full article ">Figure 2
<p>The dependencies of the component concentrations of the reaction mixture in cumene hydroperoxide decomposition versus time: dots are the experimental data, and lines are the calculation by the model; (<b>a</b>) ROOH; (<b>b</b>) ROH; (<b>c</b>) C<sub>6</sub>H<sub>5</sub>C(CH<sub>3</sub>) = CH<sub>2</sub>; (<b>d</b>) C<sub>6</sub>H<sub>5</sub>C(O)CH<sub>3</sub>; (<b>e</b>) C<sub>6</sub>H<sub>5</sub>OH.</p>
Full article ">Figure 2 Cont.
<p>The dependencies of the component concentrations of the reaction mixture in cumene hydroperoxide decomposition versus time: dots are the experimental data, and lines are the calculation by the model; (<b>a</b>) ROOH; (<b>b</b>) ROH; (<b>c</b>) C<sub>6</sub>H<sub>5</sub>C(CH<sub>3</sub>) = CH<sub>2</sub>; (<b>d</b>) C<sub>6</sub>H<sub>5</sub>C(O)CH<sub>3</sub>; (<b>e</b>) C<sub>6</sub>H<sub>5</sub>OH.</p>
Full article ">Figure 3
<p>The dependencies of cumene hydroperoxide accumulation versus time. <b><span style="color:#FF3399">Without catalyst</span><span style="color:lime">Ca</span> <span style="color:#00B0F0">Sr </span>Ba</b>. G = 0.6 L/min, T = 393 K, P = 1 atm, [RH]<sub>0</sub> = 6.19 mol/L, [ROOH]<sub>0</sub> = 0.15 mol/L, [ROH]<sub>0</sub> = 0.08 mol/L, [HCOOH]<sub>0</sub> = 0.018 mol/L, [C<sub>6</sub>H<sub>5</sub>C(O)CH<sub>3</sub>]<sub>0</sub> = 0.024 mol/L, and [Cat]<sub>0</sub> = 1 mmol/L.</p>
Full article ">Figure 4
<p>Time to reach the maximum concentration of cumene hydroperoxide depending on the temperature of cumene oxidation. <b><span style="color:#FF3399">Without catalyst</span><span style="color:lime">Ca</span> <span style="color:#00B0F0">Sr </span>Ba</b>. G = 0.6 L/min, P = 1 atm, [RH]<sub>0</sub> = 6.19 mol/L, [ROOH]<sub>0</sub> = 0.15 mol/L, [ROH]<sub>0</sub> = 0.08 mol/L, [HCOOH]<sub>0</sub> = 0.018 mol/L, [C<sub>6</sub>H<sub>5</sub>C(O)CH<sub>3</sub>]<sub>0</sub> = 0.024 mol/L, and [Cat]<sub>0</sub> = 1 mmol/L.</p>
Full article ">Figure 5
<p>The dependencies of criterion C (<b>a</b>), [ROOH]<sub>max</sub> (<b>b</b>), cumene conversion at t<sub>max</sub> (<b>c</b>), and selectivity at t<sub>max</sub> (<b>d</b>) versus cumene oxidation temperature. <b><span style="color:#FF3399">Without catalyst</span> <span style="color:lime">Ca</span> <span style="color:#00B0F0">Sr </span>Ba</b>. G = 0.6 L/min, P = 1 atm, [RH]<sub>0</sub> = 6.19 mol/L, [ROOH]<sub>0</sub> = 0.15 mol/L, [ROH]<sub>0</sub> = 0.08 mol/L, [HCOOH]<sub>0</sub> = 0.018 mol/L, [C<sub>6</sub>H<sub>5</sub>C(O)CH<sub>3</sub>]<sub>0</sub> = 0.024 mol/L, and [Cat]<sub>0</sub> = 1 mmol/L.</p>
Full article ">Figure 6
<p>The dependencies of criterion <span class="html-italic">C</span> (<b>a</b>), <span class="html-italic">t</span><sub>max</sub> (<b>b</b>), [ROOH]<sub>max</sub> (<b>c</b>), cumene conversion at <span class="html-italic">t</span><sub>max</sub> (<b>d</b>), and selectivity at <span class="html-italic">t</span><sub>max</sub> (<b>e</b>) versus initial catalyst concentration.</p>
Full article ">
12 pages, 3150 KiB  
Article
Interaction of Aggregated Cationic Porphyrins with Human Serum Albumin
by Mario Samperi, Serena Vittorio, Laura De Luca, Andrea Romeo and Luigi Monsù Scolaro
Int. J. Mol. Sci. 2023, 24(3), 2099; https://doi.org/10.3390/ijms24032099 - 20 Jan 2023
Cited by 5 | Viewed by 1833
Abstract
The interaction of an equilibrium mixture of monomeric and aggregated cationic trans-5,15-bis(N-methylpyridinium-4-yl)-10,15-bis-diphenylporphine (t-H2Pagg) chloride salt with human serum albumin (HSA) has been investigated through UV/Vis absorption, fluorescence emission, circular dichroism and resonant light scattering [...] Read more.
The interaction of an equilibrium mixture of monomeric and aggregated cationic trans-5,15-bis(N-methylpyridinium-4-yl)-10,15-bis-diphenylporphine (t-H2Pagg) chloride salt with human serum albumin (HSA) has been investigated through UV/Vis absorption, fluorescence emission, circular dichroism and resonant light scattering techniques. The spectroscopic evidence reveals that both the monomeric t-H2Pagg and its aggregates bind instantaneously to HSA, leading to the formation of a tight adduct in which the porphyrin is encapsulated within the protein scaffold (S430) and to clusters of aggregated porphyrins in electrostatic interaction with the charged biomolecules. These latter species eventually interconvert into the final S430 species following pseudo-first-order kinetics. Molecular docking simulations have been performed to get some insights into the nature of the final adduct. Analogously to hemin bound to HSA, the obtained model supports favorable interactions of the porphyrin in the same 1B subdomain of the protein. Hydrophobic and van der Waals energy terms are the main contributions to the calculated ΔGbind value of −117.24 kcal/mol. Full article
(This article belongs to the Special Issue Porphyrin and Biomolecules: A Long-Lasting Friendship)
Show Figures

Figure 1

Figure 1
<p>UV/Vis spectra of monomeric t-H<sub>2</sub>Pagg no salt added (black curve), and t-H<sub>2</sub>Pagg aggregates (red curve) obtained upon addition of NaCl. Inset: kinetic traces recorded for the decrease of monomeric t-H<sub>2</sub>Pagg at λ = 419 nm (black circles) and the formation of t-H<sub>2</sub>Pagg aggregates at λ = 450 nm (red circles). Solid lines represent the result of the global fit obtained with Equation (1) (<span class="html-italic">k</span><sub>0</sub> = 0.0036 ± 0.00097 s<sup>−1</sup>, <span class="html-italic">k<sub>c</sub></span> = 0.0237 ± 0.0007 s<sup>−1</sup>, m = 4.3 ± 0.6, <span class="html-italic">n</span> = 3.0 ± 0.5). Experimental conditions: [t-H<sub>2</sub>Pagg] = 5 μM, [NaCl] = 100 mM, phosphate buffer 1 mM, pH = 7.4, T = 298 K.</p>
Full article ">Figure 2
<p>UV/Vis spectra of t-H<sub>2</sub>Pagg aggregates (red curve) after addition of HSA at t = 0 s (light green curve) and at t = 60 min (dark green curve). Grey spectra and arrow indicate over time evolution. Inset: change over time of the extinction value at λ = 430 nm relative to the formation of the t-H<sub>2</sub>Pagg/HSA adduct. Solid line represents the fit obtained with Equation (2) (<span class="html-italic">k</span> = 0.1083 ± 0.0056 s<sup>−1</sup>, <span class="html-italic">n</span> = 0.94 ± 0.06). Experimental conditions: [t-H<sub>2</sub>Pagg] = 5 μM, [NaCl] = 100 mM, [HSA] = 100 μM, phosphate buffer 1 mM, pH = 7.4, T = 298 K.</p>
Full article ">Figure 3
<p>RLS spectra of t-H<sub>2</sub>Pagg aggregates (red curve), after addition of HSA at t = 0 s (light green curve) and at t = 60 min (dark green curve). Grey spectra and arrow indicate over time evolution. Inset: chance over time of the intensity at λ = 484 nm relative to the disassembling of t-H<sub>2</sub>Pagg aggregates. The black solid line represents the fit obtained with Equation (2) (<span class="html-italic">k</span> = 0.0996 ± 0.0046 s<sup>−1</sup>, <span class="html-italic">n</span> = 0.95 ± 0.05). Experimental conditions: [t-H<sub>2</sub>Pagg] = 5 μM, [NaCl] = 100 mM, [HSA] = 100 μM, phosphate buffer 1 mM, pH = 7.4, T = 298 K.</p>
Full article ">Figure 4
<p>CD spectra of t-H<sub>2</sub>Pagg no salt added (black curve), t-H<sub>2</sub>Pagg aggregates (red curve), after addition of HSA at t = 0 s (light green curve) and at t = 60 min (dark green curve). Experimental conditions: [tH<sub>2</sub>Pagg] = 5 μM, [NaCl] = 100 mM, [HSA] = 100 μM, phosphate buffer 1 mM, pH = 7.4, T = 298 K.</p>
Full article ">Figure 5
<p>Fluorescence emission spectra of t-H<sub>2</sub>Pagg no salt added (black curve), t-H<sub>2</sub>Pagg aggregates (red curve), after addition of HSA at t = 0 s (light green curve) and at t = 60 min (dark green curve). Experimental conditions: [t-H<sub>2</sub>Pagg] = 5 μM, [NaCl] = 100 mM, [HSA] = 100 μM, phosphate buffer 1 mM, pH = 7.4, T = 298 K, λ<sub>EX</sub> = 438 nm (the intensity of the emission spectra have been corrected for the absorbance of the samples).</p>
Full article ">Figure 6
<p>(<b>A</b>) Docking pose of t-H<sub>2</sub>Pagg with the lowest ΔG<sub>bind</sub> into the 1B subdomain of HSA (PDB ID 1N5U). The protein surface is colored according to the electrostatic potential, while t-H<sub>2</sub>Pagg is shown as cyan CPK. (<b>B</b>) Close view of the binding mode of t-H<sub>2</sub>Pagg into HSA. t-H<sub>2</sub>Pagg is represented as cyan sticks, while the residues of HSA involved in the interactions are displayed as grey sticks. H-bonds are shown as yellow dashed lines, while π-stacking and π-cation interactions are represented as cyan and green dashed lines, respectively.</p>
Full article ">Scheme 1
<p>Pictorial sketch of the molecular structure of cationic t-H<sub>2</sub>Pagg (chloride salt) and a fractal aggregate formed by increasing the ionic strength of the solution.</p>
Full article ">Scheme 2
<p>Model for the interaction of aggregated (black clusters) and monomeric (red circles) t-H<sub>2</sub>Pagg and HSA (green).</p>
Full article ">
24 pages, 5418 KiB  
Article
Interaction of Bis-(sodium-sulfopropyl)-Disulfide and Polyethylene Glycol on the Copper Electrodeposited Layer by Time-of-Flight Secondary-Ion Mass Spectrometry
by Robert Mroczka, Agnieszka Słodkowska, Agata Ładniak and Agnieszka Chrzanowska
Molecules 2023, 28(1), 433; https://doi.org/10.3390/molecules28010433 - 3 Jan 2023
Cited by 6 | Viewed by 2516
Abstract
The interactions of the functional additives SPS (bis-(sodium-sulfopropyl)-disulfide) and polyethylene glycol (PEG) in the presence of chloride ions were studied by time-of-flight secondary-ion mass spectrometry (TOF-SIMS) in combination with cyclic voltammetry measurements (CV). The PEG, thiolate, and chloride surface coverages were estimated and [...] Read more.
The interactions of the functional additives SPS (bis-(sodium-sulfopropyl)-disulfide) and polyethylene glycol (PEG) in the presence of chloride ions were studied by time-of-flight secondary-ion mass spectrometry (TOF-SIMS) in combination with cyclic voltammetry measurements (CV). The PEG, thiolate, and chloride surface coverages were estimated and discussed in terms of their electrochemical suppressing/accelerating abilities. The conformational influence of both the gauche/trans thiolate molecules, as well as around C-C and C-O of PEG, on the electrochemical properties were discussed. The contribution of the hydrophobic interaction of -CH2-CH2- of PEG with chloride ions was only slightly reduced after the addition of SPS, while the contribution of Cu-PEG adducts diminished strongly. SPS and PEG demonstrated significant synergy by significant co-adsorption. It was shown that the suppressing abilities of PEG that rely on forming stable Cu-PEG adducts, identified in the form C2H4O2Cu+ and C3H6OCu+, were significantly reduced after the addition of SPS. The major role of thiolate molecules adsorbed on a copper surface in reducing the suppressing abilities of PEG rely on the efficient capture of Cu2+ ions, diminishing the available copper ions for the ethereal oxygen of PEG. Full article
(This article belongs to the Special Issue Electrochemistry of Thin Films and Nanostructured Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Cyclic voltammetry curve recorded for the base electrolyte. (<b>b</b>) Cyclic voltammetry curves recorded for the base electrolyte and after the addition of 400 ppm of PEG8000, 30 ppm of Cl, and SPS at concentrations of 0.5, 1, 2, 5, 10, and 25 ppm. Due to the very stable value of OCP (<a href="#app1-molecules-28-00433" class="html-app">Figure S1, Supplementary Materials</a>) that equaled −151mV +/− 2 mV for all samples’ overpotential is also shown in the top axis. (<b>c</b>) CV curves shown as a function of time. The time from 0 to 30 s and the overpotential from −0.6 to 0 V (OCP) corresponds to the forward scan, while the time from 30 to 60 s and the overpotential from 0 (OCP) to −0.6 V corresponds to the reverse scan that was shown in <a href="#molecules-28-00433-f001" class="html-fig">Figure 1</a>a,b, respectively. The wire position from 0 to 1 mm corresponds to the forward scan from −0.6 V to OCP (overpotential 0 V), while the wire position from 1 to 2 mm corresponds to the reverse scan from OCP to −0.6 V.</p>
Full article ">Figure 2
<p>Exchange current densities <span class="html-italic">j</span><sub>0,1</sub> (<b>b</b>) and <span class="html-italic">j</span><sub>0,2</sub> (<b>a</b>) and the ratio <span class="html-italic">j</span><sub>0,2</sub>/<span class="html-italic">j</span><sub>0,1</sub> (<b>c</b>) for the base, base/PEG, base/PEG/Cl, and base/PEG/Cl/SPS at concentrations 0.5, 1, 2, 5, 10, and 15 ppm. The orange bars correspond to the forward scan and the green bars to the reverse scan.</p>
Full article ">Figure 3
<p>Distribution of intensity of the fragments: C<sub>2</sub>H<sub>5</sub>O<sup>+</sup>, CH<sub>2</sub>Cl<sup>+</sup>, CH<sub>2</sub>OCu<sup>+</sup>, C<sub>2</sub>H<sub>4</sub>OCu+, C<sub>3</sub>H<sub>6</sub>OCu<sup>+</sup>, C<sub>2</sub>H<sub>4</sub>O<sub>2</sub>Cu<sup>+</sup>, Cu<sub>2</sub>Cl<sup>+</sup>, the ratio of CH<sub>2</sub>Cl<sup>+</sup>/C<sub>2</sub>H<sub>5</sub>O<sup>+</sup>, C<sub>3</sub>H<sub>6</sub>OCu<sup>+</sup>/C<sub>2</sub>H<sub>5</sub>O<sup>+</sup>, and C<sub>2</sub>H<sub>4</sub>O<sub>2</sub>Cu<sup>+</sup>/C<sub>2</sub>H<sub>5</sub>O<sup>+</sup> along the wire position for the base electrolyte, base/PEG, base/PEG/Cl, and base/PEG/Cl/SPS at concentrations of 0.5, 1, 2, 5, 10, and 15 ppm. The wire position corresponds to the applied overpotential during the CV experiment.</p>
Full article ">Figure 4
<p>Distribution of the ratio of C<sub>2</sub>H<sub>5</sub>O<sup>+</sup>/Cu<sub>2</sub>Cl<sup>+</sup> as a function of the wire position for the base solution, base/PEG, base/PEG/Cl, and base/PEG/SPS at concentrations of 0.5, 1, 2, 5, 10, and 15 ppm.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>f</b>). The ratio of intensities of CH<sub>2</sub>Cl/C<sub>2</sub>H<sub>5</sub>O<sup>+</sup>, C<sub>3</sub>H<sub>6</sub>OCu<sup>+</sup>/C<sub>2</sub>H<sub>5</sub>O<sup>+</sup>, and C<sub>2</sub>H<sub>4</sub>O<sub>2</sub>Cu<sup>+</sup>/C<sub>2</sub>H<sub>5</sub>O<sup>+</sup>, obtained by dividing the ratios of the latter fragments of the adjacent sample to the ratio of the former sample. For example, the assignment SPS0.5ppm/PEG/Cl corresponds to the ratio CH<sub>2</sub>Cl<sup>+</sup>/C<sub>2</sub>H<sub>5</sub>O<sup>+</sup> for SPS 0.5 ppm (adjacent sample) divided by the ratio CH<sub>2</sub>Cl<sup>+</sup>/C<sub>2</sub>H<sub>5</sub>O<sup>+</sup> for PEG/Cl (the former sample).</p>
Full article ">Figure 6
<p>Distribution of the intensity of the fragments: CH<sub>2</sub>SO<sub>3</sub><sup>−</sup>, CuS<sup>−</sup>, SO<sub>4</sub><sup>−</sup>, C<sub>2</sub>H<sub>3</sub>SO<sub>3</sub><sup>−</sup>, CuSO<sup>−</sup>, C<sub>3</sub>H<sub>5</sub>SO<sub>3</sub><sup>−</sup>, CuSC<sub>3</sub>H<sub>6</sub>SO<sub>3</sub><sup>−</sup>, Cu<sub>2</sub>Cl<sub>3</sub><sup>−</sup>, the ratio of CH<sub>2</sub>SO<sub>3</sub><sup>−</sup>/C<sub>3</sub>H<sub>5</sub>SO<sub>3</sub><sup>−</sup>, and total thiols along with the wire position for the base electrolyte, base/PEG, base/PEG/Cl, and base/PEG/Cl/SPS at concentrations of 0.5, 1, 2, 5, 10, and 15 ppm. The wire position corresponds to the applied overpotential during the CV experiment.</p>
Full article ">Figure 7
<p>Distribution of the ratio of C<sub>2</sub>H<sub>5</sub>O<sup>+</sup>/total thiols (<b>a</b>), C<sub>2</sub>H<sub>5</sub>O<sup>−</sup>/Cu<sub>2</sub>Cl<sub>3</sub><sup>−</sup> (<b>b</b>), and total thiols/Cu<sub>2</sub>Cl<sub>3</sub><sup>−</sup> (<b>c</b>) as a function of the wire position for the base solution, base/PEG, base/PEG/Cl, and base/PEG/SPS at concentrations of 0.5, 1, 2, 5, 10, and 15 ppm.</p>
Full article ">Figure 8
<p>Possible molecular arrangements of MPS and PEG molecules on and into the copper layer during electrodeposition. Please note that some MPS molecules are incorporated into the copper deposit. The chloride adlayer was omitted for clarity.</p>
Full article ">Figure 9
<p>3D projection of the possible molecular arrangements of MPS adsorbed units in a PEG molecular surrounding.</p>
Full article ">
13 pages, 1176 KiB  
Article
Silver Dependent Enantiodivergent Gold(I) Catalysed Asymmetric Intramolecular Hydroamination of Alkenes: A Theoretical Study
by Ruchi Dixit, Himanshu Sharma, Francine Agbossou-Niedercorn, Kumar Vanka and Christophe Michon
Catalysts 2022, 12(11), 1392; https://doi.org/10.3390/catal12111392 - 8 Nov 2022
Cited by 2 | Viewed by 1874
Abstract
We report a theoretical study of the first silver-dependent enantiodivergent gold-catalysed reaction. The combination of a single chiral binuclear gold(I) chloride complex and silver perchlorate catalyses the asymmetric intramolecular hydroamination of alkenes and affords both enantiomers of the products by applying a simple [...] Read more.
We report a theoretical study of the first silver-dependent enantiodivergent gold-catalysed reaction. The combination of a single chiral binuclear gold(I) chloride complex and silver perchlorate catalyses the asymmetric intramolecular hydroamination of alkenes and affords both enantiomers of the products by applying a simple solvent change from toluene to methanol. A gold-silver chloride adduct that occurs only in methanol appears to control the enantioinversion. If one gold atom coordinates and activates the alkene moiety, the other gold is included in an adduct with silver chloride, which coordinates a methanol solvent molecule and further interacts with the amine function. If the use of toluene implies free anions and affords (S)-enantiomer, methanol allows a proximal interaction with the amine, leads to an opposite stereodifferentiation of the two diastereomeric intermediates during the final protodeauration step and results in the (R)-enantiomer. Full article
(This article belongs to the Special Issue Recent Applications of Metal Catalysts in Organic Syntheses)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The proposed reaction mechanism for the formation of the amine product (<span class="html-italic">S</span>)-<b>2</b> in toluene using catalyst <b>4a</b>. The electronic energy profile is shown at PBE0/def2-TZVP//PBE/def-TZVP level of theory. All values are in kcal/mol.</p>
Full article ">Figure 2
<p>The proposed reaction mechanism for the formation of the amine product (<span class="html-italic">R</span>)-<b>2</b> in methanol using catalyst <b>4c</b>; calculations at the PBE0/def2-TZVP//PBE/def-TZVP level of theory with DFT. All values are in kcal/mol.</p>
Full article ">Figure 3
<p>The optimised geometry of intermediate <b>3_b</b>′ implied in the reaction mechanism in methanol using catalyst <b>4c</b> and one molecule of solvent methanol. All the hydrogen atoms attached to carbon atoms are omitted for the purpose of clarity. Intermediate <b>3_b</b>′ implied in the reaction mechanism in methanol using catalyst <b>4c</b> and one molecule of solvent methanol.</p>
Full article ">Figure 4
<p>The proposed reaction mechanism for the formation of the amine product (<span class="html-italic">S</span>)-<b>2</b> in methanol using catalyst <b>4c</b>, and without the assistance of one methanol molecule. The electronic energy profile is shown at PBE0/def2-TZVP//PBE/def-TZVP level of theory. All values are in kcal/mol.</p>
Full article ">Scheme 1
<p>Enantiodivergent intramolecular hydroamination of alkenes catalysed by gold(I) cationic complex and preparation of gold(I) cationic complexes <b>4a</b>, <b>4b</b> and <b>4c</b>.</p>
Full article ">
16 pages, 2605 KiB  
Article
New Phosphonite Ligands with High Steric Demand and Low Basicity: Synthesis, Structural Properties and Cyclometalated Complexes of Pt(II)
by María M. Alcaide, Matteo Pugliesi, Eleuterio Álvarez, Joaquín López-Serrano and Riccardo Peloso
Inorganics 2022, 10(8), 109; https://doi.org/10.3390/inorganics10080109 - 29 Jul 2022
Viewed by 2971
Abstract
Two phosphonite ligands bearing the highly sterically demanding 2,6-bis (2,6-dimethylphenyl)phenyl group (ArXyl2), PArXyl2(OPhNO2)2 and PArXyl2(OPhNO2,Me)2, were prepared from the parent dihalophosphines PArXyl2 [...] Read more.
Two phosphonite ligands bearing the highly sterically demanding 2,6-bis (2,6-dimethylphenyl)phenyl group (ArXyl2), PArXyl2(OPhNO2)2 and PArXyl2(OPhNO2,Me)2, were prepared from the parent dihalophosphines PArXyl2X2 (X = Cl, Br) and the corresponding phenols, 4-nitrophenol and 4-nitro-2,6-dimethylphenol, respectively. DFT methods were used to examine their structural features and to determine three steric descriptors, namely the Tolman cone angle, the percentage of buried volume, and the percentage of the coordination sphere protected by the ligand. A comparison with the related terphenyl phosphines is also provided. Reactions of PArXyl2(OPhNO2)2 and PArXyl2(OPhNO2,Me)2 with several Pt(II) precursors were investigated, revealing a high tendency of both phosphonites to undergo C-H activation processes and generate five- or six-membered cyclometalated structures. The coordination chemistry of the new ligands was explored with isolation, among others, of three carbonyl complexes, 1-3∙CO, and the triphenylphosphine adduct 3∙PPh3. X-ray diffraction methods permitted the determination of the solid-state structures of the mononuclear methyl carbonyl complex 1∙CO, the dinuclear chloride-bridged complex 2 and the doubly cyclometalated complex 3∙SMe2, including the conformations adopted by the ligands upon coordination. All of the new compounds were characterized by multinuclear NMR spectroscopy in solution. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic views of the 3D structures of <b>PAr<sup>Xyl</sup></b><b><sup><sub>2</sub></sup>(OPh<sup>NO</sup></b><b><sup><sub>2</sub></sup>)<sub>2</sub></b> and <b>PAr<sup>Xyl</sup></b><b><sup><sub>2</sub></sup>(OPh<sup>NO</sup></b><b><sup><sub>2</sub>,Me</sup>)<sub>2</sub></b>. Curved arrows indicate fast rotations which account for the apparent symmetry observed in NMR experiments.</p>
Full article ">Figure 2
<p>Views of the calculated gas-phase structures of PAr<sup>Xyl</sup><sup><sub>2</sub></sup>(OPh<sup>NO</sup><sup><sub>2</sub></sup>)<sub>2</sub> (<b>left</b>) and PAr<sup>Xyl</sup><sup><sub>2</sub></sup>(OPh<sup>NO</sup><sup><sub>2</sub>,Me</sup>)<sub>2</sub> (<b>right</b>) (C: grey; N: purple; O: red; P: orange). Conformations C and A relate to previously reported discussions on terphenyl phosphines ligands [<a href="#B18-inorganics-10-00109" class="html-bibr">18</a>,<a href="#B19-inorganics-10-00109" class="html-bibr">19</a>,<a href="#B26-inorganics-10-00109" class="html-bibr">26</a>]. The two planes defining the dihedral angle are that containing the central ring of the terphenyl group and that containing the P—C bond and the axis of the P lone pair.</p>
Full article ">Figure 3
<p>ORTEP view of the X-ray molecular structure of complex <b>1∙CO</b>, with ellipsoids at the 50% probability level and H atoms omitted for clarity (<b>left</b>); schematic view of the B conformation adopted by the ligand <b>PAr<sup>Xyl</sup></b><b><sup><sub>2</sub></sup>(OPh<sup>NO</sup></b><b><sup><sub>2</sub></sup>)<sub>2</sub></b> (<b>right</b>) [<a href="#B18-inorganics-10-00109" class="html-bibr">18</a>,<a href="#B26-inorganics-10-00109" class="html-bibr">26</a>]. Selected bond distances (Å) and angles (°): Pt1-P1 2.2643(8), Pt1-C1 2.054(2), Pt1-C2 2.110(3), Pt1-C3 1.901(2), C3-O1 1.126(3); P1-Pt1-C1 80.86(6); C1-Pt1-C2 90.78(9); C2-Pt1-C3 88.06(1); C3-Pt1-P1 100.41(8).</p>
Full article ">Figure 4
<p>ORTEP view of the X-ray molecular structure of complex <b>2</b>, with ellipsoids at the 50% probability level and H atoms omitted for clarity (<b>left</b>); schematic view of the coordination planes in <b>2</b> with the angle they form and the binary axis <span class="html-italic">C</span><sub>2</sub> (top <b>right</b>); schematic view of the B conformation adopted by the ligand <b>PAr<sup>Xyl</sup></b><b><sup><sub>2</sub></sup>(OPh<sup>NO</sup></b><b><sup><sub>2</sub></sup>)<sub>2</sub></b> (bottom right) [<a href="#B18-inorganics-10-00109" class="html-bibr">18</a>,<a href="#B26-inorganics-10-00109" class="html-bibr">26</a>]. Selected bond distances (Å) and angles (°): Pt1-P1 2.146(1), Pt1-C1 1.974(4), Pt1-Cl1 2.410(1), Pt1-Cl1′ 2.398(1); P1-Pt1-C1 82.6(1); C1-Pt1-Cl1′ 95.4(1); Cl-Pt1-Cl′ 82.89(4); Cl1-Pt1-P1 99.10(4); Pt1-Cl1-Pt1′ 82.60(4).</p>
Full article ">Figure 5
<p>ORTEP view of the X-ray molecular structure of complex <b>3·SMe<sub>2</sub></b>, with ellipsoids at the 50% probability level and H atoms omitted for clarity (<b>left</b>); schematic view of the B conformation adopted by the ligand <b>PAr<sup>Xyl</sup></b><b><sup><sub>2</sub></sup>(OPh<sup>NO</sup></b><b><sup><sub>2</sub>,Me</sup>)<sub>2</sub></b> (<b>right</b>) [<a href="#B17-inorganics-10-00109" class="html-bibr">17</a>,<a href="#B25-inorganics-10-00109" class="html-bibr">25</a>]. Selected bond distances (Å) and angles (°): Pt1-P1 2.1526(9), Pt1-C1 2.155(3), Pt1-C2 2.120(3), Pt1-S1 2.339(1); P1-Pt1-C1 84.18(7); C1-Pt1-S1 86.74(8); S1-Pt1-C2 99.09(9); C2-Pt1-P1 89.23(8); C1-Pt1-C2 169(1).</p>
Full article ">Figure 6
<p>Schematic view along the Pt—P bond of the conformations adopted by the two cyclometalated 6-membered rings together with the values of the O-P-Pt-C dihedral angles.</p>
Full article ">Figure 7
<p><sup>1</sup>H NMR signals of the AMX spin system due to the four Pt-CH<sub>2</sub> protons in <b>3∙CO</b> (CD<sub>2</sub>Cl<sub>2</sub>) with the corresponding <sup>1</sup>H-<sup>1</sup>H (A-M), <sup>1</sup>H-<sup>31</sup>P (A-X and M-X), and <sup>1</sup>H-<sup>195</sup>Pt couplings.</p>
Full article ">Scheme 1
<p>Synthesis of the terphenyl phosphonite ligands, <b>PAr<sup>Xyl</sup></b><b><sup><sub>2</sub></sup>(OPh<sup>NO</sup></b><b><sup><sub>2</sub></sup>)<sub>2</sub></b> and <b>PAr<sup>Xyl</sup></b><b><sup><sub>2</sub></sup>(OPh<sup>NO</sup></b><b><sup><sub>2</sub>,Me</sup>)<sub>2</sub></b>.</p>
Full article ">Scheme 2
<p>Synthesis of complexes <b>1</b> and <b>2</b>.</p>
Full article ">Scheme 3
<p>Synthesis of complexes <b>3</b>.</p>
Full article ">
9 pages, 1981 KiB  
Article
Sphingolipidomics of Bovine Pink Eye: A Pilot Study
by Paul L. Wood and Lynda M. J. Miller
Vet. Sci. 2022, 9(8), 388; https://doi.org/10.3390/vetsci9080388 - 28 Jul 2022
Cited by 1 | Viewed by 2033
Abstract
Sphingolipids are essential structural components of tear film that protect the surface of the eye from dehydration. A detailed analysis of the effects of pink eye infections on the sphingolipidome in cattle has not previously been undertaken. We recently published a new assay [...] Read more.
Sphingolipids are essential structural components of tear film that protect the surface of the eye from dehydration. A detailed analysis of the effects of pink eye infections on the sphingolipidome in cattle has not previously been undertaken. We recently published a new assay utilizing high-resolution mass spectrometric monitoring of the chloride adducts of sphingolipids that provides enhanced sensitivity and specificity. Utilizing this assay, we monitored decreases in the levels of tear film ceramides with short-chain fatty acids, hydroxy-ceramides, phytoceramides, and hydroxy-phytoceramides. Dihydroceramide levels were unaltered and increased levels of ceramides with long-chain fatty acids (24:0 and 24:1) were monitored in cattle with pink eye. The data from this pilot study (n = 8 controls and 8 pink eye) demonstrate a major disruption of the lipid tear film layer in pink eye disease, that can result in severe eye irritation and damage. Full article
(This article belongs to the Section Veterinary Biomedical Sciences)
Show Figures

Figure 1

Figure 1
<p>Tear film levels of ceramides (Cer d18:1/x) in controls and cows with pink eye (P. Eye). The peak intensities of individual ceramides are expressed as a ratio of the peak intensity of ceramide d18:1/16:0. *, significant <span class="html-italic">t</span>-test statistics: 16:1 (0.013); 18:1 (0.021); 22:0 (0.027); 24:1 (0.039).</p>
Full article ">Figure 2
<p>Tear film levels of hydroxy-ceramides (Hydroxy-Cer d18:1/x) in controls and cows with pink eye (P. Eye). The peak intensities of individual hydroxy-ceramides are expressed as a ratio of the peak intensity of ceramide d18:1/16:0. *, significant <span class="html-italic">t</span>-test statistics: 16:0 (0.023); 18:0 (0.014); 18:1 (0.020); 20:1 (0.0083).</p>
Full article ">Figure 3
<p>Tear film levels of phytoceramides (Phyto-Cer) in controls and cows with pink eye (P. Eye). The peak intensities of individual phytoceramides are expressed as a ratio of the peak intensity of ceramide d18:1/16:0. *, significant <span class="html-italic">t</span>-test statistics: 34:0-O3 (0.024), 36:0-O3 (0.016), 36:1-O3 (0.01), 38:0-O3 (0.011), 40:0-O3 (0.040), 42:0-O3 (0.043), 44:0-O3 (0.036), 46:0-O3 (0.012), 46:1-O3 (0.034), 48:0-O3 (0.0086), and 48:1-O3 (0.026).</p>
Full article ">Figure 4
<p>Tear film levels of hydroxy-phytoceramides (CerX-O4) in controls and cows with pink eye (P. Eye). The peak intensities of individual hydroxy-phytoceramides are expressed as a ratio of the peak intensity of ceramide d18:1/16:0. *, significant <span class="html-italic">t</span>-test statistics: 38:0-O4 (0.0064), 40:0-O4 (0.0056), 40:1-O4 (0.0070), 42:0-O4 (0.0075), 42:1-O4 (0.0048), 44:0-O4 (0.013), 44:1-O4 (0.015), 46:0-O4 (0.0030), 48:0-O4 (0.10), and 48:1-O4 (0.018).</p>
Full article ">Figure 5
<p>Tear film levels of di-hydroxy-phytoceramides (CerX-O5) in controls and cows with pink eye (P. Eye). The peak intensities of individual di-hydroxy-phytoceramides are expressed as a ratio of the peak intensity of ceramide d18:1/16:0. *, significant <span class="html-italic">t</span>-test statistics: 42:0-O5 (0.021) and 44:0-O5 (0.0090).</p>
Full article ">Figure 6
<p>Schematic of sphingolipid metabolism. Hydroxylation of free sphingolipid bases is catalyzed by monooxygenases (SUR2) and of sphingolipid bases in dihydroceramides via hydroxylases (DES2). The introduction of hydroxylation at position 2 of the fatty acid substituents of ceramides occurs via de novo synthesis with a hydroxy fatty acid (CERS, ceramide synthase) and via lipid modification by ceramide fatty acyl 2-hydroxylase (FA2H; phytoceramide → 2-hydroxy-phytoceramide).</p>
Full article ">
8 pages, 1493 KiB  
Short Note
6-Chloro-3H-benzo[d][1,2,3]dithiazol-2-ium Chloride
by Alexander J. Nicholls and Ian R. Baxendale
Molbank 2022, 2022(1), M1339; https://doi.org/10.3390/M1339 - 11 Feb 2022
Viewed by 2131
Abstract
This short note describes the synthesis of an amorphous benzo[1,2,3]dithiazole chloride salt (commonly known as a ‘Herz salt’) by use of the Herz reaction. Hetero- and homolytic transformations of this species to a variety of useful adducts in medicinal and materials chemistry are [...] Read more.
This short note describes the synthesis of an amorphous benzo[1,2,3]dithiazole chloride salt (commonly known as a ‘Herz salt’) by use of the Herz reaction. Hetero- and homolytic transformations of this species to a variety of useful adducts in medicinal and materials chemistry are well established, although there are limited examples of isolation in the literature, and characterisation data is even harder find. While several studies have confirmed the structure of the benzodithiazole ring beyond doubt, (having generated suitably crystalline salts with large counterions for XRD-analysis), there remains value in understanding and optimising the synthesis of the simple, amorphous polymorphs. For the first time, MS data is provided for this compound and a new mechanism of its formation is proposed based upon new experimental observations and data. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The full ASAP-MS spectrum obtained from the sample of (<b>2</b>) generated in this study according to the experimental. Y-axis is (total) ion count.</p>
Full article ">Figure 2
<p>The structure of expected product (<b>2</b>) and a possible structure of the impurity detected by ASAP-MS.</p>
Full article ">Figure 3
<p>A region (185–195 Da) of the ASAP-MS spectrum obtained from the sample of (<b>2</b>) generated in this study according to the experimental. Y-axis is (total) ion count. Below are structures with their exact isotopic mass (the structure detected in this spectrum is likely <b>2b</b>).</p>
Full article ">Figure 4
<p>A region (187–197 Da) of the ASAP-MS spectrum obtained from the sample of (<b>3</b>) generated in this study according to the experimental. Y-axis is total ion count.</p>
Full article ">Figure 5
<p>A region (50–220 ppm) of the solid-state <sup>13</sup>C NMR and CPTOSS spectrum of compound <b>2</b>.</p>
Full article ">Scheme 1
<p>The Herz reaction, starting with aniline (<b>1</b>).</p>
Full article ">Scheme 2
<p>The hydrolysis of <b>2</b> to sulfoxide <b>3</b>, followed by ring-opened thiophenol <b>4</b>.</p>
Full article ">Scheme 3
<p>A proposed mechanism for the formation of title compound (<b>2</b>) by the Herz reaction.</p>
Full article ">
16 pages, 2108 KiB  
Article
Systematic Evaluation of HILIC Stationary Phases for Global Metabolomics of Human Plasma
by Farideh Hosseinkhani, Luojiao Huang, Anne-Charlotte Dubbelman, Faisa Guled, Amy C. Harms and Thomas Hankemeier
Metabolites 2022, 12(2), 165; https://doi.org/10.3390/metabo12020165 - 9 Feb 2022
Cited by 23 | Viewed by 5328
Abstract
Polar hydrophilic metabolites have been identified as important actors in many biochemical pathways. Despite continuous improvement and refinement of hydrophilic interaction liquid chromatography (HILIC) platforms, its application in global polar metabolomics has been underutilized. In this study, we aimed to systematically evaluate polar [...] Read more.
Polar hydrophilic metabolites have been identified as important actors in many biochemical pathways. Despite continuous improvement and refinement of hydrophilic interaction liquid chromatography (HILIC) platforms, its application in global polar metabolomics has been underutilized. In this study, we aimed to systematically evaluate polar stationary phases for untargeted metabolomics by using HILIC columns (neutral and zwitterionic) that have been exploited widely in targeted approaches. To do so, high-resolution mass spectrometry was applied to thoroughly investigate selectivity, repeatability and matrix effect at three pH conditions for 9 classes of polar compounds using 54 authentic standards and plasma matrix. The column performance for utilization in untargeted metabolomics was assessed using plasma samples with diverse phenotypes. Our results indicate that the ZIC-c HILIC column operated at neutral pH exhibited several advantages, including superior performance for different classes of compounds, better isomer separation, repeatability and high metabolic coverage. Regardless of the column type, the retention of inorganic ions in plasma leads to extensive adduct formation and co-elution with analytes, which results in ion-suppression as part of the overall plasma matrix effect. In ZIC-c HILIC, the sodium chloride ion effect was particularly observed for amino acids and amine classes. Successful performance of HILIC for separation of plasma samples with different phenotypes highlights this mode of separation as a valuable approach in global profiling of plasma sample and discovering the metabolic changes associated with health and disease. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overview of performance score and selectivity under chromatographic conditions. (<b>a</b>) Individual score of column performance for polar metabolites from different classes of compounds. (<b>b</b>) Column selectivity and chromatographic peak performance of representative metabolites from different classes under five different LC method conditions. Anomer mutarotation causes the glucose signal to split into two separate peaks as shown on both columns. The arrows represent the zoom figures. Blue: ZIC-c HILIC at pH 3; Red: ZIC-c HILIC at pH 7; Green: BEH Amide at pH 3; Purple: BEH Amide at pH 7; Yellow: BEH Amide at pH 10.</p>
Full article ">Figure 2
<p>Analysis of the relative contribution to the retention mechanism on different HILIC columns. (<b>a</b>) Correlation analysis on ZIC-cHILIC and (<b>b</b>) Correlation analysis on BEH-amide. Individual metabolites are shown as a: 2-hydroxybutyric acid; b: lactic acid; c: AMP; d: NAD+.</p>
Full article ">Figure 3
<p>Extracted ion chromatogram of metabolites influenced by sodium chloride on the ZIC-c HILIC and BEH-amide columns. Citrulline and glutamine showed the increased MS responses of sodium adducts in plasma and salt sample (pure sodium chloride solution) versus neat sample (no sodium chloride). Tryptophan and adenosine showed the increased MS responses of chloride adducts in plasma and salt sample (pure sodium chloride solution) versus neat sample (no sodium chloride).</p>
Full article ">Figure 4
<p>Repeatability evaluation of peak areas and retention times of 42 representative compounds during inter- and intra-batch analysis using ZIC-c HILIC and BEH-amide columns. Metabolite names with an RSD above 20% are listed on the right side of the cut-off line.</p>
Full article ">Figure 5
<p>(<b>a</b>) Principal component analysis (PCA) score plot of pre-processed untargeted features in ZIC-c and BEH-amide. Each phenotype was subjected to a triplicate sample analysis; (<b>b</b>) Retention factor (K) distribution of detected features on each column.</p>
Full article ">
13 pages, 1603 KiB  
Article
Synthesis, Spectroscopic, and Biological Assessments on Some New Rare Earth Metal Adrenaline Adducts
by Sulaiman A. Al Yousef, Asma S. Al-Wasidi, Ibtisam I. S. AlZahrani, Hotoun I. Thawibaraka, Ahmed M. Naglah, Shaima A. El-Mowafi, Omar B. Ibrahim, Moamen S. Refat and Ahmed Gaber
Crystals 2021, 11(12), 1536; https://doi.org/10.3390/cryst11121536 - 9 Dec 2021
Viewed by 2232
Abstract
Adrenaline (Adr) reacts with chlorides of Y3+, Ce3+, Nd3+ and Sm3+ in methanol at 60 °C to yield metal ion adducts of definite composition. These compounds are characterized by elemental analyses, molar conductivity, UV-Vis., 1H–NMR, Raman [...] Read more.
Adrenaline (Adr) reacts with chlorides of Y3+, Ce3+, Nd3+ and Sm3+ in methanol at 60 °C to yield metal ion adducts of definite composition. These compounds are characterized by elemental analyses, molar conductivity, UV-Vis., 1H–NMR, Raman laser, scanning electron microscopy, energy-dispersive X-ray spectroscopy (EDX), and mid infrared spectral measurement investigations. The adducts are found to have the formulae [Y2(Adr)2(H2O)8]Cl3.8H2O, [Ce(Adr)2(H2O)2]Cl3.10H2O, [Nd(Adr)2(H2O)2]Cl3.6H2O, and [Sm(Adr)2(H2O)2]Cl3.12H2O, respectively. The two phenolic groups of the catechol moiety are linked to central metal ions based on the infrared and Raman laser spectra. The new compounds were tested against five gram-positive and two-gram negative bacteria, in addition to two Aspergillus strains. Metal adducts were shown to have stronger antibacterial and antifungal properties than free adrenaline compounds. Full article
(This article belongs to the Special Issue Research about Vital Organic Chelates and Metal Ion Complexes)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>): Suggested structures of [Y<sub>2</sub>(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>8</sub>]Cl<sub>3</sub>·8H<sub>2</sub>O adduct. (<b>b</b>): Suggested structures of [Ce(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·10H<sub>2</sub>O, [Nd(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·6H<sub>2</sub>O and [Sm(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·12H<sub>2</sub>O adducts.</p>
Full article ">Figure 2
<p>Infrared spectra of [Y<sub>2</sub>(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>8</sub>]Cl<sub>3</sub>·8H<sub>2</sub>O, [Ce(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·10H<sub>2</sub>O, [Nd(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·6H<sub>2</sub>O, and [Sm(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·12H<sub>2</sub>O adducts.</p>
Full article ">Figure 3
<p>Raman spectra of [Y<sub>2</sub>(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>8</sub>]Cl<sub>3</sub>·8H<sub>2</sub>O, [Ce(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·10H<sub>2</sub>O, [Nd(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·6H<sub>2</sub>O, and [Sm(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·12H<sub>2</sub>O adducts.</p>
Full article ">Figure 4
<p>Electronic spectrum of adrenaline free and their complexes [Y<sub>2</sub>(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>8</sub>]Cl<sub>3</sub>·8H<sub>2</sub>O, and [Ce(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·10H<sub>2</sub>O, [Nd(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·6H<sub>2</sub>O, Sm(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·12H<sub>2</sub>O.</p>
Full article ">Figure 5
<p>(<b>A</b>). EDX free adrenaline spectrum. (<b>B</b>). EDX spectrum of [Y<sub>2</sub>(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>8</sub>]Cl<sub>3</sub>·8H<sub>2</sub>O adduct. (<b>C</b>). EDX spectrum of [Ce(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·10H<sub>2</sub>O adduct. (<b>D</b>). EDX spectrum of [Nd(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·6H<sub>2</sub>O adduct. (<b>E</b>). EDX spectrum of [Sm(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·12H<sub>2</sub>O adduct.</p>
Full article ">Figure 5 Cont.
<p>(<b>A</b>). EDX free adrenaline spectrum. (<b>B</b>). EDX spectrum of [Y<sub>2</sub>(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>8</sub>]Cl<sub>3</sub>·8H<sub>2</sub>O adduct. (<b>C</b>). EDX spectrum of [Ce(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·10H<sub>2</sub>O adduct. (<b>D</b>). EDX spectrum of [Nd(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·6H<sub>2</sub>O adduct. (<b>E</b>). EDX spectrum of [Sm(Adr)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]Cl<sub>3</sub>·12H<sub>2</sub>O adduct.</p>
Full article ">
11 pages, 5229 KiB  
Communication
2-Pyridylselenenyl versus 2-Pyridyltellurenyl Halides: Symmetrical Chalcogen Bonding in the Solid State and Reactivity towards Nitriles
by Ivan V. Buslov, Alexander S. Novikov, Victor N. Khrustalev, Mariya V. Grudova, Alexey S. Kubasov, Zhanna V. Matsulevich, Alexander V. Borisov, Julia M. Lukiyanova, Maria M. Grishina, Anatoly A. Kirichuk, Tatiyana V. Serebryanskaya, Andreii S. Kritchenkov and Alexander G. Tskhovrebov
Symmetry 2021, 13(12), 2350; https://doi.org/10.3390/sym13122350 - 7 Dec 2021
Cited by 19 | Viewed by 3642
Abstract
The synthesis of 2-pyridyltellurenyl bromide via Br2 oxidative cleavage of the Te–Te bond of dipyridylditelluride is reported. Single-crystal X-ray diffraction analysis of 2-pyridyltellurenyl bromide demonstrated that the Te atom of 2-pyridyltellurenyl bromide was involved in four different noncovalent contacts: Te⋯Te interactions, two [...] Read more.
The synthesis of 2-pyridyltellurenyl bromide via Br2 oxidative cleavage of the Te–Te bond of dipyridylditelluride is reported. Single-crystal X-ray diffraction analysis of 2-pyridyltellurenyl bromide demonstrated that the Te atom of 2-pyridyltellurenyl bromide was involved in four different noncovalent contacts: Te⋯Te interactions, two Te⋯Br ChB, and one Te⋯N ChB contact forming 3D supramolecular symmetrical framework. In contrast to 2-pyridylselenenyl halides, the Te congener does not react with nitriles furnishing cyclization products. 2-Pyridylselenenyl chloride was demonstrated to easily form the corresponding adduct with benzonitrile. The cyclization product was studied by the single-crystal X-ray diffraction analysis, which revealed that in contrast to earlier studied cationic 1,2,4-selenadiazoles, here we observed that the adduct with benzonitrile formed supramolecular dimers via Se⋯Se interactions in the solid state, which were never observed before for 1,2,4-selenadiazoles. Full article
(This article belongs to the Section Chemistry: Symmetry/Asymmetry)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of adducts of 2-pyridylselenenyl halides and nitriles showing the position of the Se-centered σ-holes and the N atom lone pair.</p>
Full article ">Figure 2
<p>Ball-and-stick representations of the crystal structures of <b>16</b>, demonstrating symmetrical Te⋯Br and Te⋯N chalcogen bonding and Te⋯Te interactions. Grey and light-grey, blue, dark-green, and brown spheres represent carbon, hydrogen, nitrogen, tellurium, and bromine atoms, respectively.</p>
Full article ">Figure 3
<p>Ball-and-stick representations of the crystal structures of <b>1</b>, demonstrating symmetrical Se⋯Cl and Se⋯N chalcogen bonding. Grey and light-grey spheres represent carbon and hydrogen atoms, respectively.</p>
Full article ">Figure 4
<p>Ball-and-stick representations of the crystal structures of <b>15</b>, demonstrating symmetrical Se⋯Cl chalcogen bonding and Se⋯Se interactions. Grey and light-grey spheres represent carbon and hydrogen, respectively.</p>
Full article ">Figure 5
<p>Ball-and-stick representations of the crystal structures of <b>6</b>–<b>12</b>, demonstrating supramolecular dimerization via ChB or HB. Grey and light-grey spheres represent carbon and hydrogen, respectively.</p>
Full article ">Figure 6
<p>Laplacian of electron density distribution ∇<sup>2</sup>ρ(<b>r</b>) (<b>left</b> panels), visualization of electron localization function (ELF, <b>center</b> panels) and reduced density gradient (RDG, <b>right</b> panels) analyses for noncovalent interactions Se⋯Cl in the crystal structure of <b>1</b>. Bond critical points (3, −1) are shown in blue, the color scale for the ELF and RDG maps is presented in a.u., length units–Å.</p>
Full article ">Figure 7
<p>Laplacians of electron density distribution ∇<sup>2</sup>ρ(<b>r</b>) (<b>left</b> panels), visualization of electron localization function (ELF, <b>center</b> panels) and reduced density gradient (RDG, <b>right</b> panels) analyses for noncovalent interactions Te⋯Br and Te⋯Te in the X-ray structure <b>16</b>. Bond critical points (3, −1) are shown in blue, the color scale for the ELF and RDG maps is presented in a.u., length units–Å.</p>
Full article ">Figure 8
<p>Laplacians of electron density distribution ∇<sup>2</sup>ρ(<b>r</b>) (<b>left</b> panels), visualization of electron localization function (ELF, <b>center</b> panels) and reduced density gradient (RDG, <b>right</b> panels) analyses for noncovalent interactions Se⋯Se, Se⋯Cl, and Cl⋯N in the X-ray structure <b>15</b>. Bond critical points (3, −1) are shown in blue, the color scale for the ELF and RDG maps is presented in a.u., length units–Å.</p>
Full article ">Scheme 1
<p>Synthesis of <b>3</b>–<b>15</b>.</p>
Full article ">
12 pages, 5952 KiB  
Article
Hydrogen-Bonded and Halogen-Bonded: Orthogonal Interactions for the Chloride Anion of a Pyrazolium Salt
by Steven van Terwingen, Daniel Brüx, Ruimin Wang and Ulli Englert
Molecules 2021, 26(13), 3982; https://doi.org/10.3390/molecules26133982 - 29 Jun 2021
Cited by 9 | Viewed by 2981
Abstract
In the hydrochloride of a pyrazolyl-substituted acetylacetone, the chloride anion is hydrogen-bonded to the protonated pyrazolyl moiety. Equimolar co-crystallization with tetrafluorodiiodobenzene (TFDIB) leads to a supramolecular aggregate in which TFDIB is situated on a crystallographic center of inversion. The iodine atom in the [...] Read more.
In the hydrochloride of a pyrazolyl-substituted acetylacetone, the chloride anion is hydrogen-bonded to the protonated pyrazolyl moiety. Equimolar co-crystallization with tetrafluorodiiodobenzene (TFDIB) leads to a supramolecular aggregate in which TFDIB is situated on a crystallographic center of inversion. The iodine atom in the asymmetric unit acts as halogen bond donor, and the chloride acceptor approaches the σ-hole of this TFDIB iodine subtending an almost linear halogen bond, with Cl···I = 3.1653(11) Å and Cl···I–C = 179.32(6)°. This contact is roughly orthogonal to the N–H···Cl hydrogen bond. An analysis of the electron density according to Bader’s Quantum Theory of Atoms in Molecules confirms bond critical points (bcps) for both short contacts, with ρbcp = 0.129 for the halogen and 0.321eÅ3 for the hydrogen bond. Our halogen-bonded adduct represents the prototype for a future class of co-crystals with tunable electron density distribution about the σ-hole contact. Full article
Show Figures

Figure 1

Figure 1
<p>The halogen X exhibits an electron deficient site (red, <math display="inline"><semantics> <mrow> <mrow> <mi mathvariant="normal">δ</mi> </mrow> <mo>+</mo> </mrow> </semantics></math>) in direction of the <math display="inline"><semantics> <mi mathvariant="sans-serif">σ</mi> </semantics></math>-bond to R. The nucleophile Y can interact with this positively charged region <span class="html-italic">via</span> its lone pair, thus forming a halogen bond.</p>
Full article ">Figure 2
<p>Chemical structures of heterobifunctional molecules utilized by our group for crystal engineering exhibiting nitrogen lone pairs.</p>
Full article ">Figure 3
<p>Displacement ellipsoid plot [<a href="#B30-molecules-26-03982" class="html-bibr">30</a>] of one molecule in the asymmetric residue of <b>1</b> (80% probability, C bonded hydrogens omitted). Selected intramolecular distances and angles (Å, °): O1–C2 1.311(2), O2–C4 1.278(2), C2–C3 1.395(3), C3–C4 1.424(3), C2–C3–C8–C9 85.7(2), N1–N2–C11–C16 28.6(2).</p>
Full article ">Figure 4
<p>Difference Fourier contour map [<a href="#B30-molecules-26-03982" class="html-bibr">30</a>] of the acetylacetone moiety in both molecules contained in the asymmetric residue of <b>1</b>. Contour lines are drawn at <math display="inline"><semantics> <mrow> <mn>0.2</mn> </mrow> </semantics></math> <span class="html-italic">e</span><math display="inline"><semantics> <msup> <mi mathvariant="sans-serif">Å</mi> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math> (red: positive difference, green: negative difference, blue: zero lines).</p>
Full article ">Figure 5
<p>Overlay plot [<a href="#B30-molecules-26-03982" class="html-bibr">30</a>] of both molecules in the asymmetric residue in <b>1</b> (black: molecule 1 as shown in <a href="#molecules-26-03982-f003" class="html-fig">Figure 3</a>, red: molecule 2 under symmetry operator <math display="inline"><semantics> <mrow> <mo>−</mo> <mi>x</mi> <mo>,</mo> <mo>−</mo> <mi>y</mi> <mo>,</mo> <mo>−</mo> <mi>z</mi> </mrow> </semantics></math>; C bonded hydrogens omitted).</p>
Full article ">Figure 6
<p>Displacement ellipsoid plot [<a href="#B30-molecules-26-03982" class="html-bibr">30</a>] of <b>2</b> (80% probability, C bonded hydrogens omitted). Selected intramolecular distances and angles (Å, °): I1···Cl1 3.1653(11), Cl1···N1 2.970(2), I1···Cl1···N1 73.99(4), C17–I1···Cl1 179.32(6), Cl1···H1N–N1 172(3). Symmetry operation: a = <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>−</mo> <mi>x</mi> <mo>,</mo> <mo>−</mo> <mi>y</mi> <mo>,</mo> <mn>1</mn> <mo>−</mo> <mi>z</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Hirshfeld surface [<a href="#B46-molecules-26-03982" class="html-bibr">46</a>] around Cl1 mapped with <math display="inline"><semantics> <msub> <mi>d</mi> <mi>norm</mi> </msub> </semantics></math>; regions marked in red represent directions of short, those in blue of long contact distances.</p>
Full article ">Figure 8
<p>Atomic basins [<a href="#B47-molecules-26-03982" class="html-bibr">47</a>,<a href="#B48-molecules-26-03982" class="html-bibr">48</a>] in <b>2</b>; intramolecular bond paths and the conventional hydrogen bond are shown as solid black lines, the halogen bond and non-classical hydrogen bonds as dashed black lines.</p>
Full article ">Figure 9
<p>Laplacian of the electron density in <b>2</b>; positive values in blue, negative values in red, contours at <math display="inline"><semantics> <mrow> <mo>±</mo> <msup> <mn>2</mn> <mi>n</mi> </msup> <mspace width="0.166667em"/> <mi>·</mi> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math> a. u. (<math display="inline"><semantics> <mrow> <mn>0</mn> <mo>≤</mo> <mi>n</mi> <mo>≤</mo> <mn>20</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 10
<p>Electrostatic potential for <b>2</b>, mapped on an isosurface of electron density <math display="inline"><semantics> <mi>ρ</mi> </semantics></math> = 0.05 a. u. [<a href="#B48-molecules-26-03982" class="html-bibr">48</a>]; red areas are associated with a positive value (0.480 a. u.), cyan areas with negative values (−0.0675 a. u.) and green areas with an ESP (0.115 a. u.).</p>
Full article ">
13 pages, 1626 KiB  
Article
Human Brain Lipidomics: Utilities of Chloride Adducts in Flow Injection Analysis
by Paul L. Wood, Kathleen A. Hauther, Jon H. Scarborough, Dustin J. Craney, Beatrix Dudzik, John E. Cebak and Randall L. Woltjer
Life 2021, 11(5), 403; https://doi.org/10.3390/life11050403 - 28 Apr 2021
Cited by 9 | Viewed by 2337
Abstract
Ceramides have been implicated in a number of disease processes. However, current means of evaluation with flow infusion analysis (FIA) have been limited primarily due to poor sensitivity within our high-resolution mass spectrometry lipidomics analytical platform. To circumvent this deficiency, we investigated the [...] Read more.
Ceramides have been implicated in a number of disease processes. However, current means of evaluation with flow infusion analysis (FIA) have been limited primarily due to poor sensitivity within our high-resolution mass spectrometry lipidomics analytical platform. To circumvent this deficiency, we investigated the potential of chloride adducts as an alternative method to improve sensitivity with electrospray ionization. Chloride adducts of ceramides and ceramide subfamilies provided 2- to 50-fold increases in sensitivity both with analytical standards and biological samples. Chloride adducts of a number of other lipids with reactive hydroxy groups were also enhanced. For example, monogalactosyl diacylglycerols (MGDGs), extracted from frontal lobe cortical gray and subcortical white matter of cognitively intact subjects, were not detected as ammonium adducts but were readily detected as chloride adducts. Hydroxy lipids demonstrate a high level of specificity in that phosphoglycerols and phosphoinositols do not form chloride adducts. In the case of choline glycerophospholipids, the fatty acid substituents of these lipids could be monitored by MS2 of the chloride adducts. Monitoring the chloride adducts of a number of key lipids offers enhanced sensitivity and specificity with FIA. In the case of glycerophosphocholines, the chloride adducts also allow determination of fatty acid substituents. The chloride adducts of lipids possessing electrophilic hydrogens of hydroxyl groups provide significant increases in sensitivity. In the case of glycerophosphocholines, chloride attachment to the quaternary ammonium group generates a dominant anion, which provides the identities of the fatty acid substituents under MS2 conditions. Full article
(This article belongs to the Special Issue Multi-Omics for the Understanding of Brain Diseases)
Show Figures

Figure 1

Figure 1
<p>Ceramide families monitored in human frontal cortex gray matter (GM) and subcortical white matter (WM). These lipid families include ceramides (Cer), hydroxyceramides (OH-Cer), hexosylceramides (Hex-Cer), hexosylhydroxylceramides (Hex-OH-Cer), phytoceramides (Phy-Cer), and ceramide-phosphoethanolamines (Cer-PE). Relative levels (endogenous lipid peak intensity/peak intensity of a stable isotope internal standard) were corrected for wet weight differences. The internal standard for these determinations was 2 nmoles of [<sup>13</sup>C<sub>16</sub>]Cer d18:1/16:0. The specific masses utilized are summarized in <a href="#life-11-00403-t001" class="html-table">Table 1</a>. Mean ± SEM (N = 12).</p>
Full article ">Figure 2
<p>MS<sup>2</sup> spectrum for MGDG 34:1, clearly demonstrating that this is MGDG 16:1/18:0. Theoretical 16:1 = 253.21730 (0.76 ppm) and 18:0 = 283.2643 (0.61 ppm), also see <a href="#life-11-00403-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 3
<p>Monoacylglycerols (MG), diacylglycerols (DG), and monogalactosyl diacylglycerol 34:1 (MGDG) found in human frontal cortex gray matter (GM) and white matter (WM). The MS<sup>2</sup> spectrum for MGDG 34:1 is presented in <a href="#life-11-00403-f003" class="html-fig">Figure 3</a>. Relative levels (endogenous lipid peak intensity/peak intensity of a stable isotope internal standard) were corrected for wet weight differences. The internal standard used for these determinations was 2 nmoles of [<sup>13</sup>C<sub>3</sub>]DG 36:2. The specific masses utilized are summarized in <a href="#life-11-00403-t002" class="html-table">Table 2</a>. Mean ± SEM (N = 12).</p>
Full article ">Figure 4
<p>Sphingomyelins (SM) and hydroxysphingomyelins (OH-SM) found in human frontal cortical gray matter (GM) and subcortical white matter (WM). Relative levels (endogenous lipid peak intensity/peak intensity of a stable isotope internal standard) were corrected for wet weight differences. The internal standard used for these determinations was 10 nmoles of [<sup>2</sup>H<sub>7</sub>]SM d18:1/16:0. The specific masses utilized are summarized in <a href="#life-11-00403-t003" class="html-table">Table 3</a>. Mean ± SEM (N = 12).</p>
Full article ">Figure 5
<p>Phosphatidylcholines (PC) found in human frontal cortical gray matter (GM) and subcortical white matter (WM). Relative levels (endogenous lipid peak intensity/peak intensity of a stable isotope internal standard) were corrected for wet weight differences. The internal standard used for these determinations was 10 nanomoles of [<sup>13</sup>C<sub>40</sub>]PC 32:0. The specific masses utilized are summarized in <a href="#life-11-00403-t004" class="html-table">Table 4</a>. Mean ± SEM (N = 12).</p>
Full article ">
Back to TopTop