Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 15 de 15
Filter
Add more filters










Publication year range
1.
Polymers (Basel) ; 15(20)2023 Oct 19.
Article in English | MEDLINE | ID: mdl-37896387

ABSTRACT

A new set of microporous organic polymers (POPs) containing diphosphine derivatives synthesized by knitting via Friedel-Crafts has been attained. These amorphous three-dimensional materials have been prepared by utilizing diphosphines, 1,3,5-triphenylbenzene, and biphenyl as nucleophile aromatic groups, dimethoxymethane as the electrophilic linker, and FeCl3 as a promoting catalyst. These polymer networks display moderate thermal stability and high microporosity, boasting BET surface areas above 760 m2/g. They are capable of coordinating with palladium acetate, using the phosphine derivative as an anchoring center, and have proven to be highly efficient catalysts in Suzuki-Miyaura coupling reactions involving bromo- and chloroarenes under environmentally friendly (using water and ethanol as solvents) and aerobic conditions. These supported catalysts have achieved excellent turnover numbers (TON) and turnover frequencies (TOF), while maintaining good recyclability without significant loss of activity or Pd leaching after five consecutive reaction cycles.

2.
Int J Mol Sci ; 24(3)2023 Feb 01.
Article in English | MEDLINE | ID: mdl-36769053

ABSTRACT

By employing a variation of the polyamidation method using in situ silylated diamines and acid chlorides, it was possible to obtain a rod-type polyamide: poly(p-phenylene terephthalamide) (PPTA, a polymer used in the high-value-added material Kevlar), with a molecular weight much higher than that obtained with the classical and industrial polyamidation method. The optimization of the method has consisted of using, together with the silylating agent, a mixture of pyridine and a high-pKa tertiary amine. The research was complemented by a combination of nuclear magnetic resonance and molecular simulation studies, which determined that the improvements in molecular weight derive mainly from the formation of silylamide groups in the growing polymer.


Subject(s)
Nylons , Polymers , Molecular Weight , Magnetic Resonance Spectroscopy , Diamines
3.
J Am Chem Soc ; 145(1): 527-536, 2023 01 11.
Article in English | MEDLINE | ID: mdl-36542758

ABSTRACT

The synthesis of aryl-alkynyl compounds is usually achieved via Sonogashira catalysis, but this is inefficient for fluorinated aryls. An alternative method reported by Shirakawa and Hiyama, using alkynylstannanes and hemilabile PN ligands, works apparently fine for conventional aryls, but it is also poor for fluorinated aryls. The revision of the unusual literature cycle reveals the existence and nature of unreported byproducts and uncovers coexisting cycles and other aspects that explain the reasons for the conflict. This knowledge provides a full understanding of the real complexity of these aryl/alkynylstannane systems and the deviations of their evolution from that of a classic Stille process, providing the clues to design several very efficient alternatives for the catalytic synthesis of the desired ArF-alkynyl compounds in almost quantitative yield. The same protocols are also very efficient for the catalytic synthesis of alkynyl-alkynyl' hetero- and homocoupling.


Subject(s)
Tin Compounds , Catalysis
4.
Inorg Chem ; 59(24): 18287-18294, 2020 Dec 21.
Article in English | MEDLINE | ID: mdl-33289542

ABSTRACT

The NiII literature complex cis-[Ni(C6F5)2(THF)2] is a synthon of cis-Ni(C6F5)2 that allows us to establish a protocol to measure and compare the ligand effect on the NiII → Ni0 reductive elimination step (coupling), often critical in catalytic processes. Several ligands of different types were submitted to this Ni-meter comparison: bipyridines, chelating diphosphines, monodentate phosphines, PR2(biaryl) phosphines, and PEWO ligands (phosphines with one potentially chelate electron-withdrawing olefin). Extremely different C6F5-C6F5 coupling rates, ranging from totally inactive (producing stable complexes at room temperature) to those inducing almost instantaneous coupling at 25 °C, were found for the different ligands tested. The PR2(biaryl) ligands, very efficient for coupling in Pd, are slow and inefficient in Ni, and the reason for this difference is examined. In contrast, PEWO type ligands are amazingly efficient and provide the lowest coupling barriers ever observed for NiII complexes; they yield up to 96% C6F5-C6F5 coupling in 5 min at 25 °C (the rest is C6F5H) and 100% coupling with no hydrolysis in 8 h at -22 to -53 °C.

5.
Dalton Trans ; 49(32): 11336-11345, 2020 Aug 18.
Article in English | MEDLINE | ID: mdl-32766614

ABSTRACT

Neutral palladium(ii) complexes [Pd(Rf)X(P-L)] (Rf = 3,5-C6Cl2F3, X = Cl, I, OTf) with P-P (dppe and dppf) and P-N (PPh2(bzN)) ligands have chelated structures in the solid-state, except for P-L = dppf and X = Cl, were chelated and dimeric bridged structures are found. The species present in solution in different solvents (CDCl3, THF, NMP and HMPA) have been characterised by 19F and 31P{1H} NMR and conductivity studies. Some [Pd(Rf)X(P-L)] complexes are involved in equilibria with [Pd(Rf)(solv)(P-L)]X, depending on the solvent and X. The ΔH° and ΔS° values of these equilibria explain the variations of ionic vs. neutral complexes in the range 183-293 K. Overall the order of coordination strength of solvents and anionic ligands is: HMPA ≫ NMP > THF and I-, Cl- > TfO-. This coordination preference is determining the complexes participating in the alkynyl transmetalation from PhC[triple bond, length as m-dash]CSnBu3 to [Pd(Rf)X(P-L)] (X = OTf, I) in THF and subsequent coupling. Very different reaction rates and stability of intermediates are observed for similar complexes, revealing neglected complexities that catalytic cycles have to deal with. Rich information on the evolution of these Stille systems after transmetalation has been obtained that leads to proposal of a common behaviour for complexes with dppe and PPh2(bzN), but a different evolution for the complexes with dppf: this difference leads the latter to produce PhC[triple bond, length as m-dash]CRf and black Pd, whereas the two former yield PhC[triple bond, length as m-dash]CRf and [Pd(C[triple bond, length as m-dash]CPh)(SnBu3)(dppe)] or [Pd(C[triple bond, length as m-dash]CPh)(SnBu3){PPh2(bzN)}].

6.
Dalton Trans ; 49(23): 7852-7861, 2020 Jun 21.
Article in English | MEDLINE | ID: mdl-32463416

ABSTRACT

This work describes the synthesis of the gold(i) complexes of phosphine sulphides. The formation of these new derivatives has been confirmed by X-ray crystallography. The coordination of gold(i) with the sulphur atom of the phosphine sulphides favors the inhibition of topoisomerase I as well as a high cytotoxicity of the gold(i)-complexed compounds against the cancer line A549 with IC50 values in the nanomolar range and IC50 values below 5 µM against the SKOV3 cell line. It should be noted that the cytotoxicities observed for the new gold(i) complexes are higher than those observed for phosphine sulphide ligands before binding to gold. Furthermore, the results also indicate that the presence of a nitrogenated heterocycle, such as tetrahydroquinoline or quinoline, is also necessary for the TopI inhibition to be maintained. In addition, no toxicity was observed when the non-cancerous lung fibroblast cell line (MRC5) was treated with the new phosphine sulphide gold(i) complexes prepared.


Subject(s)
Antineoplastic Agents/pharmacology , Coordination Complexes/pharmacology , DNA Topoisomerases, Type I/metabolism , Topoisomerase I Inhibitors/pharmacology , A549 Cells , Antineoplastic Agents/chemical synthesis , Antineoplastic Agents/chemistry , Cell Line, Tumor , Cell Proliferation/drug effects , Coordination Complexes/chemical synthesis , Coordination Complexes/chemistry , Crystallography, X-Ray , Density Functional Theory , Dose-Response Relationship, Drug , Drug Screening Assays, Antitumor , Gold/chemistry , Gold/pharmacology , Humans , Models, Molecular , Molecular Structure , Phosphines/chemistry , Phosphines/pharmacology , Structure-Activity Relationship , Sulfides/chemistry , Sulfides/pharmacology , Topoisomerase I Inhibitors/chemical synthesis , Topoisomerase I Inhibitors/chemistry
7.
Chem Commun (Camb) ; 54(83): 11809-11812, 2018 Oct 16.
Article in English | MEDLINE | ID: mdl-30280158

ABSTRACT

Chiral-diamine catalyzed addition of ZnMe2 to PhC(O)CF2X (in dichloromethane at -30 °C) affords fluorinated alkyl tertiary alcohols in high yield (quantitative for X = H, F, Cl; 84% for X = CF3) and up to 99% ee. These conditions are similarly very efficient for other various ArC(O)CF3 molecules. A fine analysis of the results can be performed based on a double-cycle mechanism.

8.
ACS Appl Mater Interfaces ; 10(31): 26195-26205, 2018 Aug 08.
Article in English | MEDLINE | ID: mdl-30001102

ABSTRACT

A new generation of porous polymer networks has been obtained in quantitative yield by reacting two rigid trifunctional aromatic monomers (1,3,5-triphenylbenzene and triptycene) with two ketones having electron-withdrawing groups (trifluoroacetophenone and isatin) in superacidic media. The resulting amorphous networks are microporous materials, with moderate Brunauer-Emmett-Teller surface areas (from 580 to 790 m2 g-1), and have high thermal stability. In particular, isatin yields networks with a very high narrow microporosity contribution, 82% for triptycene and 64% for 1,3,5-triphenylbenzene. The existence of favorable interactions between lactams and CO2 molecules has been stated. The materials show excellent CO2 uptakes (up to 207 mg g-1 at 0 °C/1 bar) and can be regenerated by vacuum, without heating. Under postcombustion conditions, their CO2/N2 selectivities are comparable to those of other organic porous networks. Because of the easily scalable synthetic method and their favorable characteristics, these materials are very promising as industrial adsorbents.

9.
Dalton Trans ; 46(25): 8083-8090, 2017 Jun 27.
Article in English | MEDLINE | ID: mdl-28604880

ABSTRACT

A polymer [Pd(CH2SO2C6H4Me)2]n is obtained by thermolysis of cis-[Pd(CH2SO2C6H4Me)2(NCMe)2] to release the MeCN ligands. The corresponding coordination sites are then occupied by weak Pd-O bonds, easier to break than the previous Pd-N bonds. This allows us to produce from the polymer cis complexes containing ligands weaker than NCMe, such as acetone or water. The complexes cis-[Pd(CH2SO2C6H4Me)2{OC(CD3)2}2], cis-[Pd(CH2SO2C6H4Me)2(OH2)2], and cis-[Pd(CH2SO2C6H4Me)2(OH2){OC(CD3)2}], and cyclic dimers [Pd(CH2SO2C6H4Me)2(OH2)]2 with bridging methylsulphone groups are formed. The Pd : PPh3 : OH2 1 : 1 : 1 reaction of the polymer produces cis-[Pd(CH2SO2C6H4Me)2(OH2)(PPh3)], which isomerizes to trans-[Pd(CH2SO2C6H4Me)2(OH2)(PPh3)], with water O-coordinated to Pd and making hydrogen bonds to the two SO2 groups as seen in its X-ray structure. A similar role is played by RNH2 groups in the structures of trans-[Pd(CH2SO2C6H4Me)2(NH3)(PPh3)] and the dimer µ-(N2H4)(trans-[Pd(CH2SO2C6H4Me)2(PPh3)])2. In addition to these interesting intramolecular hydrogen bonding properties provided by the SO2 groups, the structural and 1H NMR data available suggest that the CH2SO2C6H4Me group is an interesting kind of strong alkyl σ donor, with high trans influence, and forms very stable Pd complexes extraordinarily resistant to reductive elimination and to hydrolysis by water at room temperature.

10.
Angew Chem Int Ed Engl ; 55(42): 13276-13280, 2016 10 10.
Article in English | MEDLINE | ID: mdl-27634585

ABSTRACT

A Pd complex, cis-[Pd(C6 F5 )2 (THF)2 ] (1), is proposed as a useful touchstone for direct and simple experimental measurement of the relative ability of ancillary ligands to induce C-C coupling. Interestingly, 1 is also a good alternative to other precatalysts used to produce Pd0 L. Complex 1 ranks the coupling ability of some popular ligands in the order Pt Bu3 >o-TolPEWO-F≈tBuXPhos>P(C6 F5 )3 ≈PhPEWO-F>P(o-Tol)3 ≈THF≈tBuBrettPhos≫Xantphos≈PhPEWO-H≫PPh3 according to their initial coupling rates, whereas their efficiency, depending on competitive hydrolysis, is ranked tBuXPhos≈Pt Bu3 ≈o-TolPEWO-F>PhPEWO-F>P(C6 F5 )3 ≫tBuBrettPhos>THF≈P(o-Tol)3 >Xantphos>PhPEWO-H≫PPh3 . This "meter" also detects some other possible virtues or complications of ligands such as tBuXPhos or tBuBrettPhos.

11.
Chemistry ; 20(45): 14800-6, 2014 Nov 03.
Article in English | MEDLINE | ID: mdl-25223511

ABSTRACT

NMR spectroscopic studies of the catalytic addition reaction of ZnEt2 to PhC(O)CF3 in the presence of three very efficient catalysts [TMEDA, tBuBOX, and L; where L is a chiral diamine synthesized from optically pure (R,R)-1,2-diphenylethylenediamine and (S)-2,2'-bis-(bromomethyl)-1,1'-binaphthalene] reveal large differences in their behavior. For the ligands TMEDA and tBuBOX, the catalysis shows no unusual features and proceeds via [(NN)Zn(Et){OC(CF3 )(Et)Ph}]. For NNL, the observation of autocatalytic asymmetric enhancement during the catalysis, and unusual inverse concentration dependence on the reaction rate, indicate the participation of an additional novel catalytic cycle that goes through a dinuclear intermediate containing one ZnEt2 and one ZnEt fragment connected by NN and OR bridges. Interestingly, the (19) F NMR signals of the main product of the reaction ([Zn(Et){OC*(CF3 )(Et)Ph}]2 ) allowed us to assess the enantioselectivity of the processes in situ without the assistance of chiral chromatography.


Subject(s)
Acetophenones/chemistry , Chlorofluorocarbons, Methane/chemistry , Diamines/chemistry , Organometallic Compounds/chemistry , Catalysis , Magnetic Resonance Spectroscopy , Stereoisomerism
12.
Dalton Trans ; 42(40): 14576-82, 2013 Oct 28.
Article in English | MEDLINE | ID: mdl-23982452

ABSTRACT

Chiral 1,2-ethylenediamines have been previously reported as active catalysts in the enantioselective addition reactions of ZnR2 to either methyl- or trifluoromethyl-ketones. Subtle changes in the molecular structure of different catalysts are described herein and lead to a dramatic effect in their catalytic activity. From these findings, we demonstrate the selective reactivity of the ligands used in the addition of ZnR2 (R = Me, Et) to methyl- and trifluoromethyl-ketones offering an enantioselective access either to chiral non-fluorinated alcohols or to chiral fluorinated tertiary alcohols. Considering the importance of the chiral trifluoromethyl carbinol fragment in several biologically active compounds, we have extended the scope of the addition reaction of ZnEt2 to several trifluoromethylketones catalyzed by (R,R)-1,2-diphenylethylenediamine derivatives. This work explores a homogeneous approach that provides excellent yields and very high ee and the use of a heterogenized tail-tied ligand affording moderate ee, high yields and allowing an easier handling and recycling.


Subject(s)
Alcohols/chemical synthesis , Diamines/chemistry , Ketones/chemistry , Organometallic Compounds/chemistry , Zinc/chemistry , Alcohols/chemistry , Molecular Structure , Organometallic Compounds/chemical synthesis , Stereoisomerism
13.
Chemistry ; 19(11): 3702-9, 2013 Mar 11.
Article in English | MEDLINE | ID: mdl-23371865

ABSTRACT

The X-ray structure of the title compound [Pd(Fmes)2 (tmeda)] (Fmes=2,4,6-tris(trifluoromethyl)phenyl; tmeda=N,N,N',N'-tetramethylethylenediamine) shows the existence of uncommon CH⋅⋅⋅FC hydrogen-bond interactions between methyl groups of the TMEDA ligand and ortho-CF3 groups of the Fmes ligand. The (19) F NMR spectra in CD2 Cl2 at very low temperature (157 K) detect restricted rotation for the two ortho-CF3 groups involved in hydrogen bonding, which might suggest that the hydrogen bond is responsible for this hindrance to rotation. However, a theoretical study of the hydrogen-bond energy shows that it is too weak (about 7 kJ mol(-1) ) to account for the rotational barrier observed (ΔH(≠) =26.8 kJ mol(-1) ), and it is the steric hindrance associated with the puckering of the TMEDA ligand that should be held responsible for most of the rotational barrier. At higher temperatures the rotation becomes fast, which requires that the hydrogen bond is continuously being split up and restored and exists only intermittently, following the pulse of the conformational changes of TMEDA.


Subject(s)
Coordination Complexes/chemistry , Organometallic Compounds/chemistry , Coordination Complexes/chemical synthesis , Crystallography, X-Ray , Hydrogen Bonding , Magnetic Resonance Spectroscopy , Models, Molecular , Molecular Conformation , Organometallic Compounds/chemical synthesis , Solutions
14.
Inorg Chem ; 46(3): 1027-32, 2007 Feb 05.
Article in English | MEDLINE | ID: mdl-17257047

ABSTRACT

When trans-[NiRf2L2] (Rf = 3,5-C6Cl2F3; L = group 15 soft monodentate weak ligand such as SbPh3 or AsPh3) is dissolved in wet (CD3)2CO, isomerization (to give cis-[NiRf2L2]) and subsequent substitutions of L by (CD3)2CO or by water occur, and several complexes containing acetone and aqua ligands are formed. The isomerization takes place in a few seconds at room temperature. The substitution reactions on the cis isomer formed are faster. The kinetics of the equilibria between all of the participating species have been studied by 19F exchange spectroscopy experiments at 217 K, and the exchange rates and rate constants have been calculated. These data reflect the weakness of acetone compared to water and AsPh3. The data obtained are the first available for square-planar nickel(II) aquacomplexes. The bulkier AsCyPh2 ligand slows down the exchange processes while the displacement of AsMePh2 is clearly disfavored. Activation entropy studies support an associative ligand substitution. All of these data fit well with the previously reported relative activity of these complexes as catalysts in norbornene polymerization.

15.
Inorg Chem ; 35(5): 1250-1256, 1996 Feb 28.
Article in English | MEDLINE | ID: mdl-11666315

ABSTRACT

The dihydrido-olefin complex OsH(2)(eta(2)-CH(2)=CHEt)(CO)(P(i)Pr(3))(2) (2) reacts with H(2)SiPh(2) to give OsH(3)(SiHPh(2))(CO)(P(i)Pr(3))(2) (3). The molecular structure of 3 has been determined by X-ray diffraction (monoclinic, space group P2(1)/c with a = 16.375(2) Å, b = 11.670(1) Å, c =18.806(2) Å, beta = 107.67(1) degrees, and Z = 4) together with ab initio calculations on the model compound OsH(3)(SiH(3))(CO)(PH(3))(2). The coordination geometry around the osmium center can be rationalized as a heavily distorted pentagonal bipyramid with one hydrido ligand and the carbonyl group in the axial positions. The two other hydrido ligands lie in the equatorial plane, one between the phosphine ligands and the other between the SiHPh(2) group and one of the phosphine ligands. Complex 3 can also be prepared by reaction of OsH(eta(2)-H(2)BH(2))(CO)(P(i)Pr(3))(2) (4) with H(2)SiPh(2). Similarly, the treatment of 4 with HSiPh(3) affords OsH(3)(SiPh(3))(CO)(P(i)Pr(3))(2) (5), while the addition of H(3)SiPh to 4 in methanol yields OsH(3){Si(OMe)(2)Ph}(CO)(P(i)Pr(3))(2) (6). Complex 2 also reacts with HGeR(3) and HSnR(3) to give OsH(3)(GeR(3))(CO)(P(i)Pr(3))(2) (GeR(3) = GeHPh(2) (7), GePh(3) (8), GeEt(3) (9)) and OsH(3)(SnR(3))(CO)(P(i)Pr(3))(2) (R = Ph (10), (n)Bu (11)), respectively. In solution, compounds 3 and 5-11 are fluxional and display similar (1)H and (31)P{(1)H} NMR spectra, suggesting that they possess a similar arrangement of ligands around the osmium atom.

SELECTION OF CITATIONS
SEARCH DETAIL
...