Herbert Plenio

Find an error

Name:
Organization: Technische Universität Darmstadt
Department: Anorganische Chemie im Zintl-Institute
Title:

TOPICS

Co-reporter:O. Halter;H. Plenio
Chemical Communications 2017 vol. 53(Issue 92) pp:12461-12464
Publication Date(Web):2017/11/16
DOI:10.1039/C7CC07018G
Fluorescence resonance energy transfer (FRET) in appropriately tagged NHC–gold complexes can be utilized to in situ observe the formation of digold species, which are known to be catalytically relevant intermediates in various gold catalyzed reactions.
Co-reporter:Roman Savka, Sabine Foro and Herbert Plenio  
Dalton Transactions 2016 vol. 45(Issue 27) pp:11015-11024
Publication Date(Web):17 Jun 2016
DOI:10.1039/C6DT01724J
The reaction of 1-amino,4-hydroxy-pentiptycene with diacetyl or acenaphthene-1,2-dione gave the respective diimines, followed by alkylation of the hydroxyl groups, and cyclization of the alkylated diimines to the respective bispentiptycene-imidazolium salts NHC·HCl. The azolium salts, being precursors to N-heterocyclic carbenes, were converted into metal complexes [(NHC)MX] (MX = CuI, AgCl, AuCl) and [(NHC)IrCl(cod)] and [(NHC)IrCl(CO)2] in good yields. In the solid state [(NHC)AgCl] displays a bowl-shaped structure of the ligand with the metal center buried within the concave unit.
Co-reporter:Dr. Roman Savka;M.Sc. Marvin Bergmann;Yuki Kanai;Sabine Foro;Dr. Herbert Plenio
Chemistry - A European Journal 2016 Volume 22( Issue 28) pp:9667-9675
Publication Date(Web):
DOI:10.1002/chem.201601474

Abstract

Based on 1-amino-4-hydroxy-triptycene, new saturated and unsaturated triptycene-NHC (N-heterocyclic carbene) ligands were synthesized from glyoxal-derived diimines. The respective carbenes were converted into metal complexes [(NHC)MX] (M=Cu, Ag, Au; X=Cl, Br) and [(NHC)MCl(cod)] (M=Rh, Ir; cod=1,5-cyclooctadiene) in good yields. The new azolium salts and metal complexes suffer from limited solubility in common organic solvents. Consequently, the introduction of solubilizing groups (such as 2-ethylhexyl or 1-hexyl by O-alkylation) is essential to render the complexes soluble. The triptycene unit infers special steric properties onto the metal complexes that enable the steric shielding of selected areas close to the metal center. Next, chiral and meso-triptycene based N-heterocyclic carbene ligands were prepared. The key step in the synthesis of the chiral ligand is the Buchwald–Hartwig amination of 1-bromo-4-butoxy-triptycene with (1S,2S)-1,2-diphenyl-1,2-diaminoethane, followed by cyclization to the azolinium salt with HC(OEt)3. The analogous reaction with meso-1,2-diphenyl-1,2-diaminoethane provides the respective meso-azolinium salt. Both the chiral and meso-azolinium salts were converted into metal complexes including [(NHC)AuCl], [(NHC)RhCl(cod)], [(NHC)IrCl(cod)], and [(NHC)PdCl(allyl)]. An in situ prepared chiral copper complex was tested in the enantioselective borylation of α,β-unsaturated esters and found to give an excellent enantiomeric ratio (er close to 90:10).

Co-reporter:Roman Vasiuta ; Herbert Plenio
Chemistry - A European Journal 2016 Volume 22( Issue 18) pp:6353-6360
Publication Date(Web):
DOI:10.1002/chem.201600264

Abstract

The Pd-catalyzed reactions of 3-chloro-bodipy with R2PH (R=Ph, Cy) provide nonfluorescent bodipy–phosphines 3-PR2–bodipy 3 a (R=Ph) and 3 b (R=Cy; quantum yield Φ<0.001). Metal complexes such as [AgCl(3 b)] and [AuCl(3 b)] were prepared and shown to display much higher fluorescence (Φ=0.073 and 0.096). In the gold complexes, the level of fluorescence was found to be qualitatively correlated with the electron density at gold. Consequently, the fluorescence brightness of [AuCl(3 b)] increases when the chloro ligand is replaced by a weakly coordinating anion, whereas upon formation of the electron-rich complex [Au(SR)(3 b)] the fluorescence is almost quenched. Related reactions of [AuCl(3 b)] with [Ag]ONf)] (Nf= nonaflate) and phenyl acetylenes enable the tracking of initial steps in gold-catalyzed reactions by using fluorescence spectroscopy. Treatment of [AuCl(3 b)] with [Ag(ONf)] gave the respective [Au(ONf)(3 b)] only when employing more than 2.5 equivalents of silver salt. The reaction of the “cationic” gold complex with phenyl acetylenes leads to the formation of the respective dinuclear cationic [{(3 b)Au}2(CCPh)]+ and an increase in the level of fluorescence. The rate of the reaction of [Au(ONf)(3 b)] with PhCCH depends on the amount of silver salt in the reaction mixture; a large excess of silver salt accelerates this transformation. In situ fluorescence spectroscopy thus provides valuable information on the association of gold complexes with acetylenes.

Co-reporter:R. Savka and H. Plenio  
Dalton Transactions 2015 vol. 44(Issue 3) pp:891-893
Publication Date(Web):12 Nov 2014
DOI:10.1039/C4DT03449J
The reactions of [MCl(cod)]2 (M = Rh, Ir) with different NHC·HX (X = Cl, I), K2CO3 in technical grade acetone under air provide simple access to various [(NHC)MX(cod)] complexes; a facile one-pot synthesis of [(NHC)MCl(CO)2] (M = Rh, Ir) is also reported.
Co-reporter:M.Sc. Pavlo Kos ;Dr. Herbert Plenio
Angewandte Chemie International Edition 2015 Volume 54( Issue 45) pp:13293-13296
Publication Date(Web):
DOI:10.1002/anie.201506918

Abstract

A Crabtree-type IrI complex tagged with a fluorescent dye (bodipy) was synthesized. The oxidative addition of H2 converts the weakly fluorescent IrI complex (Φ=0.038) into a highly fluorescent IrIII species (Φ=0.51). This fluorogenic reaction can be utilized for the detection of H2 and to probe the oxidative addition step in the catalytic hydrogenation of olefins.

Co-reporter:M.Sc. Pavlo Kos ;Dr. Herbert Plenio
Angewandte Chemie 2015 Volume 127( Issue 45) pp:13491-13494
Publication Date(Web):
DOI:10.1002/ange.201506918

Abstract

A Crabtree-type IrI complex tagged with a fluorescent dye (bodipy) was synthesized. The oxidative addition of H2 converts the weakly fluorescent IrI complex (Φ=0.038) into a highly fluorescent IrIII species (Φ=0.51). This fluorogenic reaction can be utilized for the detection of H2 and to probe the oxidative addition step in the catalytic hydrogenation of olefins.

Co-reporter:Pavlo Kos ;Dr. Herbert Plenio
Chemistry - A European Journal 2015 Volume 21( Issue 3) pp:1088-1095
Publication Date(Web):
DOI:10.1002/chem.201405316

Abstract

Several metal complexes with a boron dipyrromethene (BODIPY)-functionalized N-heterocyclic carbene (NHC) ligand 4 were synthesized. The fluorescence in [(4)(SIMes)RuCl2(ind)] complex is quenched (Φ=0.003), it is weak in [(4)PdI2(Clpy)] (Φ=0.033), and strong in [(4)AuI] (Φ=0.70). The BODIPY-tagged complexes can experience pronounced changes in the brightness of the fluorophore upon ligand-exchange and ligand-dissociation reactions. Complexes [(4)MX(1,5-cyclooctadiene)] (M=Rh, Ir; X=Cl, I; Φ=0.008–0.016) are converted into strongly fluorescent complexes [(4)MX(CO)2] (Φ=0.53–0.70) upon reaction with carbon monoxide. The unquenching of the Rh and Ir complexes appears to be a consequence of the decreased electron density at Rh or Ir in the carbonyl complexes. In contrast, the substitution of an iodo ligand in [(4)AuI] by an electron-rich thiolate decreases the brightness of the BODIPY fluorophore, rendering the BODIPY as a highly sensitive probe for changes in the coordination sphere of the transition metal.

Co-reporter:Meike Egert;Sebastian Walther
European Journal of Organic Chemistry 2014 Volume 2014( Issue 20) pp:4362-4369
Publication Date(Web):
DOI:10.1002/ejoc.201402364

Abstract

The present work establishes a new synthetic route that leads to substituted azolium salts. The base stable 1-(4-bromo-2,6-diisopropylphenyl)-3-(2,6-diisopropylphenyl)imidazolidine and 1,3-bis(4-bromo-2,6-diisopropylphenyl)imidazolidine were synthesized and the 4-Br substituents converted into various functional groups through Br/Li exchange or Pd-catalyzed cross-coupling reactions (Suzuki, Sonogashira, and vinylation). The substituted imidazolidines were oxidized, by using chloranil or N-bromosuccinimide, to provide the respective azolium salts, which are convenient precursors to N-heterocyclic carbenes.

Co-reporter:Roman Savka
European Journal of Inorganic Chemistry 2014 Volume 2014( Issue 36) pp:6246-6253
Publication Date(Web):
DOI:10.1002/ejic.201402830

Abstract

1,3,5-Tricycloalkylbenzene (cycloalkyl = C5H9, C6H11) was converted into the respective anilines (by means of nitration and reduction) and then into the corresponding diimines (with glyoxal), the cyclization of which with (HCHO)n/ZnCl2 provided the respective 1,3-bis(2,4,6-tricyclopentylphenyl)imidazolium salt in modest yields. An analogous reaction sequence that employed acenaphthene-1,2-dione instead of glyoxal yielded the two azolium salts in good yields, which were converted into the respective N-heterocyclic carbene (NHC) complexes [(NHC)AgCl], [(NHC)AuCl], [(NHC)RhCl(cod)] (cod = cyclooctadiene), and [(NHC)RhCl(CO)2].

Co-reporter:Stefanie Wolf and Herbert Plenio  
Green Chemistry 2013 vol. 15(Issue 2) pp:315-319
Publication Date(Web):21 Dec 2012
DOI:10.1039/C2GC36417D
End-of-life tire granulates were used in olefin metathesis reactions with ethene (ethenolysis); utilizing 0.0074 mmol of a Grubbs type ruthenium complex for 1.0 gram of tire granules leads to the isolation of ca. 0.5 g of organic solubles, which are primarily composed of oligomeric 1,4-cis isoprenes.
Co-reporter:Pavlo Kos;Roman Savka
Advanced Synthesis & Catalysis 2013 Volume 355( Issue 2-3) pp:439-447
Publication Date(Web):
DOI:10.1002/adsc.201200956

Abstract

Four 1-(4-R-phenoxy)-2-ethenylbenzenes (R=NMe2, H, Cl, NO2) 4a, 4b, 4c and 4d were reacted with the ruthenium complexes [RuCl2(NHC)(3-phenylindenylidene)(py)] in the presence of a protic resin to result in the formation of the respective Hoveyda-type complexes 5a–d {NHC=SIMes [1,3-bis(2,4,6-trimethylphenylimidazolin)-2-ylidene]} and 6a–d {NHC=SIPr [1,3-bis(2,6-diisopropylphenylimidazolin)-2-ylidene]} in 66–84% yield. The lower steric bulk and the decreased donation of the diaryl ether oxygen atoms in complexes 5 and 6 led to rapidly initiating precatalysts. The Ru(II/III) redox potentials of complexes 6 were determined (6a–d: ΔE=0.89–1.08 V). In the crystal structure of 5b two independent molecules were observed in the unit cell, displaying RuO distances of 226.6(4) and 230.5(3) pm. The catalytic performance of complexes 5 and 6 in various ring-closing metathesis (RCM) reactions was studied. Catalyst loadings of between 15–200 ppm are sufficient for the formation of >90% yield of the respective cyclic products. Complex 6b catalyzes the formation of N-protected 2,5-dihydropyrroles with up to TON 64,000 and TOF 256,000 h−1, of the N-protected 1,2,3,6- tetrahydropyridines with up to TON 18,200 and TOF 73,000 h−1 and of the N-protected 2,3,6,7-tetrahydroazepines with up to TON 8,100 and TOF 32,000 h−1 with yields ranging between 77 and 96%.

Co-reporter:Dr. Vasco Thiel;Dr. Klaus-Jürgen Wannowius;Dipl.-Ing. Christiane Wolff;Dr. Christina M. Thiele;Dr. Herbert Plenio
Chemistry - A European Journal 2013 Volume 19( Issue 48) pp:16403-16414
Publication Date(Web):
DOI:10.1002/chem.201204150

Abstract

Conversion–time data were recorded for various ring-closing metathesis (RCM) reactions that lead to five- or six-membered cyclic olefins by using different precatalysts of the Hoveyda type. Slowly activated precatalysts were found to produce more RCM product than rapidly activated complexes, but this comes at the price of slower product formation. A kinetic model for the analysis of the conversion–time data was derived, which is based on the conversion of the precatalyst (Pcat) into the active species (Acat), with the rate constant kact, followed by two parallel reactions: 1) the catalytic reaction, which utilizes Acat to convert reactants into products, with the rate kcat, and 2) the conversion of Acat into the inactive species (Dcat), with the rate kdec. The calculations employ two experimental parameters: the concentration of the substrate (c(S)) at a given time and the rate of substrate conversion (−dc(S)/dt). This provides a direct measure of the concentration of Acat and enables the calculation of the pseudo-first-order rate constants kact, kcat, and kdec and of kS (for the RCM conversion of the respective substrate by Acat). Most of the RCM reactions studied with different precatalysts are characterized by fast kcat rates and by the kdec value being greater than the kact value, which leads to quasistationarity for Acat. The active species formed during the activation step was shown to be the same, regardless of the nature of different Pcats. The decomposition of Acat occurs along two parallel pathways, a unimolecular (or pseudo-first-order) reaction and a bimolecular reaction involving two ruthenium complexes. Electron-deficient precatalysts display higher rates of catalyst deactivation than their electron-rich relatives. Slowly initiating Pcats act as a reservoir, by generating small stationary concentrations of Acat. Based on this, it can be understood why the use of different precatalysts results in different substrate conversions in olefin metathesis reactions.

Co-reporter:M.Sc. Roman Savka;Sabine Foro;Dr. Markus Gallei;Dr. Matthias Rehahn;Dr. Herbert Plenio
Chemistry - A European Journal 2013 Volume 19( Issue 32) pp:10655-10662
Publication Date(Web):
DOI:10.1002/chem.201300868

Abstract

Eight new N-Hoveyda-type complexes were synthesized in yields of 67–92 % through reaction of [RuCl2(NHC)(Ind)(py)] (NHC=1,3-bis(2,4,6-trimethylphenylimidazolin)-2-ylidene (SIMes) or 1,3-bis(2,6-diisopropylphenylimidazolin)-2-ylidene (SIPr), Ind=3-phenylindenylid-1-ene, py=pyridine) with various 1- or 1,2-substituted ferrocene compounds with vinyl and amine or imine substituents. The redox potentials of the respective complexes were determined; in all complexes an iron-centered oxidation reaction occurs at potentials close to E=+0.5 V. The crystal structures of the reduced and of the respective oxidized Hoveyda-type complexes were determined and show that the oxidation of the ferrocene unit has little effect on the ruthenium environment. Two of the eight new complexes were found to be switchable catalysts, in that the reduced form is inactive in the ring-opening metathesis polymerization of cis-cyclooctene (COE), whereas the oxidized complexes produce polyCOE. The other complexes are not switchable catalysts and are either inactive or active in both reduced and oxidized states.

Co-reporter:Roman D. Savka, Herbert Plenio
Journal of Organometallic Chemistry 2012 710() pp: 68-74
Publication Date(Web):
DOI:10.1016/j.jorganchem.2012.03.015
Co-reporter:Dr. Lars H. Peeck;Roman D. Savka; Herbert Plenio
Chemistry - A European Journal 2012 Volume 18( Issue 40) pp:12845-12853
Publication Date(Web):
DOI:10.1002/chem.201201010

Abstract

Reactions of the Grubbs 3rd generation complexes [RuCl2(NHC)(Ind)(Py)] (N-heterocyclic carbene (NHC)=1,3-bis(2,4,6-trimethylphenylimidazolin)-2-ylidene (SIMes), 1,3-bis(2,6-diisopropylphenylimidazolin)-2-ylidene (SIPr), or 1,3-bis(2,6-diisopropylphenylimidazol)-2-ylidene (IPr); Ind=3-phenylindenylid-1-ene, Py=pyridine) with 2-ethenyl-N-alkylaniline (alkyl=Me, Et) result in the formation of the new N-Grubbs–Hoveyda-type complexes 5 (NHC=SIMes, alkyl=Me), 6 (SIMes, Et), 7 (IPr, Me), 8 (SIPr, Me), and 9 (SIPr, Et) with N-chelating benzylidene ligands in yields of 50–75 %. Compared to their respective, conventional, O-Grubbs–Hoveyda complexes, the new complexes are characterized by fast catalyst activation, which translates into fast and efficient ring-closing metathesis (RCM) reactivity. Catalyst loadings of 15–150 ppm (0.0015–0.015 mol %) are sufficient for the conversion of a wide range of diolefinic substrates into the respective RCM products after 15 min at 50 °C in toluene; compounds 8 and 9 are the most catalytically active complexes. The use of complex 8 in RCM reactions enables the formation of N-protected 2,5-dihydropyrroles with turnover numbers (TONs) of up to 58 000 and turnover frequencies (TOFs) of up to 232 000 h−1; the use of the N-protected 1,2,3,6-tetrahydropyridines proceeds with TONs of up to 37 000 and TOFs of up to 147 000 h−1; and the use of the N-protected 2,3,6,7-tetrahydroazepines proceeds with TONs of up to 19 000 and TOFs of up to 76 000 h−1, with yields for these reactions ranging from 83–92 %.

Co-reporter:Marc Schilz and Herbert Plenio
The Journal of Organic Chemistry 2012 Volume 77(Issue 6) pp:2798-2807
Publication Date(Web):March 6, 2012
DOI:10.1021/jo202644g
The conversion–time data for 168 different Pd/Cu-catalyzed Sonogashira cross-coupling reactions of five arylacetylenes (phenylacetylene; 1-ethynyl-2-ethylbenzene; 1-ethynyl-2,4,6-R3-benzene (R = Me, Et, i-Pr)) and Me3SiCCH with seven aryl bromides (three 2-R-bromobenzenes (R = Me, Et, i-Pr); 2,6-Me2-bromobenzene and three 2,4,6-R3-bromobenzenes (R = Me, Et, i-Pr)) with four different phosphines (P-t-Bu3, t-Bu2PCy, t-BuPCy2, PCy3) were determined using quantitative gas chromatography. The stereoelectronic properties of the substituents in the aryl bromides, acetylenes, and phosphines were correlated with the performance in Sonogashira reactions. It was found that the nature of the most active Pd/PR3 complex for a Sonogashira transformation is primarily determined by the steric bulk of the acetylene; ideal catalysts are: Pd/P-t-Bu3 or Pd/t-Bu2PCy for sterically undemanding phenylacetylene, Pd/t-BuPCy2 for 2- and 2,6-substituted arylacetylenes or Me3SiCCH and Pd/PCy3 for extremely bulky acetylenes and aryl bromides. Electron-rich and sterically demanding aryl bromides with substituents in the 2- or the 2,6-position require larger amounts of catalyst than 4-substituted aryl bromides. The synthesis of tolanes with bulky groups at one of the two aryl rings is best done by placing the steric bulk at the arylacetylene, which is also the best place for electron-withdrawing substituents.
Co-reporter:Vasco Thiel ; Marina Hendann ; Klaus-Jürgen Wannowius
Journal of the American Chemical Society 2011 Volume 134(Issue 2) pp:1104-1114
Publication Date(Web):December 21, 2011
DOI:10.1021/ja208967h
Grubbs–Hoveyda-type complexes with variable 4-R (complexes 1: 4-R = NEt2, OiPr, H, F, NO2) and 5-R substituents (complexes 2: 5-R = NEt2, OiPr, Me, F, NO2) at the 2-isopropoxy benzylidene ether ligand and with variable 4-R substituents (complexes 3: 4-R = H, NO2) at the 2-methoxy benzylidene ether ligand were synthesized and the respective Ru(II/III) redox potentials (ranging from ΔE = +0.46 to +1.04 V), and UV–vis spectra recorded. The initiation kinetics of complexes 1–3 with the olefins diethyl diallyl malonate (DEDAM), butyl vinyl ether (BuVE), 1-hexene, styrene, and 3,3-dimethylbut-1-ene were investigated using UV–vis spectroscopy. Electron-withdrawing groups at the benzylidene ether ligands were found to increase the initiation rates, while electron-donating groups lead to slower precatalyst activation; accordingly with DEDAM, the complex 1(NO2) initiates almost 100 times faster than 1(NEt2). The 4-R substituents (para to the benzylidene carbon) were found to have a stronger influence on physical and kinetic properties of complexes 1 and 2 than that of 5-R groups para to the ether oxygen. The DEDAM-induced initiation reactions of complexes 1 and 2 are classified as two-step reactions with an element of reversibility. The hyperbolic fit of the kobs vs [DEDAM] plots is interpreted according to a dissociative mechanism (D). Kinetic studies employing BuVE showed that the initiation reactions simultaneously follow two different mechanistic pathways, since the kobs vs [olefin] plots are best fitted to kobs = kD·k4/k–D·[olefin]/(1 + k4/k–D·[olefin]) + kI·[olefin]. The kI·[olefin] term dominates the initiation behavior of the sterically less demanding complexes 3 and was shown to correspond to an interchange mechanism with associative mode of activation (Ia), leading to very fast precatalyst activation at high olefin concentrations. Equilibrium and rate constants for the reactions of complexes 1–3 with the bulky PCy3 were determined. In general, sterically demanding olefins (DEDAM, styrene) and Grubbs–Hoveyda type complexes 1 and 2 preferentially initiate according to the dissociative pathway; for the less bulky olefins (BuVE, 1-hexene) and complexes 1 and 2 both D and Ia are important. Activation parameters for BuVE reactions and complexes 1(NEt2), 1(H), and 1(NO2) were determined, and ΔS‡ was found to be negative (ΔS‡ = −113 to −167 J·K–1·mol–1) providing additional support for the Ia catalyst activation.
Co-reporter:Stefanie Wolf and Herbert Plenio  
Green Chemistry 2011 vol. 13(Issue 8) pp:2008-2012
Publication Date(Web):28 Jun 2011
DOI:10.1039/C1GC15265C
We report here the ethenolysis of squalene and natural rubber utilizing (NHC)(NHCewg)RuCl2(= CRR’) and Grubbs–Hoveyda complexes. 0.01 mol% [Ru] per double bond are sufficient for extensive squalene cleavage, resulting in the formation of numerous terminal olefins, which were identified by GC/MS. The depolymerization of natural rubber requires 0.1 mol% [Ru] and leads to the formation of various oligomeric isoprenes, several of which (n = 2–6) were isolated and characterized.
Co-reporter:Christoph A. Fleckenstein and Herbert Plenio  
Chemical Society Reviews 2010 vol. 39(Issue 2) pp:694-711
Publication Date(Web):16 Oct 2009
DOI:10.1039/B903646F
The strong electron-donation and the steric bulk of trialkylphosphines renders them as very useful ligands for palladium-catalyzed cross coupling reactions. This critical review reports on the synthesis of two families of trialkylphosphines (diadamantylalkylphosphines, fluorenyldialkylphosphines) and the properties of the respective palladium complexes in various cross coupling reactions, which evolved as alternatives to the classical Pd/PtBu3 system. In contrast to the latter phosphine the new classes of ligands are characterized by a highly flexible ligand design, which allows the fine tuning of catalytic properties to the specific needs of certain substrates and also enables the attachment of additional tags to impart certain useful properties onto the respective phosphines (179 references).
Co-reporter:Jan Pschierer, Natalie Peschek and Herbert Plenio  
Green Chemistry 2010 vol. 12(Issue 4) pp:636-642
Publication Date(Web):08 Feb 2010
DOI:10.1039/B924772F
We, the named authors, hereby wholly retract this Green Chemistry article. Signed: Jan Pschierer, Natalie Peschek and Herbert Plenio, Germany, September 2010. Retraction endorsed by Sarah Ruthven, Editor. Retraction published 15th September 2010.
Co-reporter:Sutapa Roy
Advanced Synthesis & Catalysis 2010 Volume 352( Issue 6) pp:1014-1022
Publication Date(Web):
DOI:10.1002/adsc.200900886

Abstract

The reactions of the N,N′-diarylimidazolium and N,N′-diarylimidazolinium salts with chlorosulfonic acid result in the formation of the respective disulfonated N-heterocyclic carbene (NHC) precursors in reasonable yields (46–77%). Water-soluble palladium catalyst complexes, in situ obtained from the respective sulfonated imidazolinium salt, sodium tetrachloropalladate (Na2PdCl4) and potassium hydroxide (KOH) in water, were successfully applied in the copper-free Sonogashira coupling reaction in isopropyl alcohol/water mixtures using 0.2 mol% catalyst loading. The preformed (disulfonatedNHC)PdCl(cinnamyl) complex was used in aqueous Suzuki–Miyaura reactions at 0.1 mol% catalyst loading. The coupling protocol reported here is very useful for Sonogashira reactions of N- and S-heterocyclic aryl bromides and chlorides with aryl- and alkylacetylenes.

Co-reporter:Jan Pschierer
European Journal of Organic Chemistry 2010 Volume 2010( Issue 15) pp:2934-2937
Publication Date(Web):
DOI:10.1002/ejoc.201000251

Abstract

The palladium complex (0.5 mol-%) of a water-soluble sulfonated fluorenylphosphane (cataCXium Fsulf) enables thefacile Suzuki–Miyaura coupling of various (heterocyclic) aryl tosylates and aryl mesylate with various (heterocyclic) boronic acids in excellent yields (> 95 %) using water as the reaction solvent.

Retraction: The following article from the European Journal of Organic Chemistry, “Suzuki–Miyaura Coupling of Aryl Tosylates and Mesylates in Water”, published online on April 15, 2010 in Wiley Online Library (www.onlinelibrary.wiley.com, doi: 10.1002/ejoc.201000251) and in print (Eur. J. Org. Chem.2010, 2934–2937), has been retracted by agreement between the corresponding author, the journal Editor, Dr. Haymo Ross, and Wiley-VCH. The retraction has been agreed because several 1H and 13C NMR spectroscopic data listed in the manuscript are incorrect, and the original mass spectra cannot be located. Attempts to repeat the synthesis of several representative products under the conditions reported in this manuscript have failed.

Co-reporter:Jan Pschierer
European Journal of Organic Chemistry 2010 Volume 2010( Issue 29) pp:
Publication Date(Web):
DOI:10.1002/ejoc.201001176

Abstract

Retraction: No Abstract

Co-reporter:Tim Vorfalt Dipl.-Ing.;Klaus-Jürgen Wannowius Dr. Dr.
Angewandte Chemie 2010 Volume 122( Issue 32) pp:5665-5668
Publication Date(Web):
DOI:10.1002/ange.201000581
Co-reporter:Jan Pschierer Dr.
Angewandte Chemie 2010 Volume 122( Issue 35) pp:6361-6364
Publication Date(Web):
DOI:10.1002/ange.201002045
Co-reporter:Lars H. Peeck and Herbert Plenio
Organometallics 2010 Volume 29(Issue 12) pp:2761-2766
Publication Date(Web):June 3, 2010
DOI:10.1021/om1002717
[(NHC)RuCl2(3-phenylindenylid-1-ene)(py)] (1) serves as a convenient starting material for the synthesis of [(NHC)(NHCewg)RuCl2(3-phenylindenylid-1-ene)] complexes 3a−3g utilizing [AgI(NHCewg)] complexes (2) as NHC transfer reagents. The respective complexes 3 display excellent activities in RCM reactions leading to tetrasubstituted olefins. The most active precatalyst, 3f, is characterized by 3,4-dichloro and N,N′-diethyl substituents and can be obtained in 94% isolated yield. The redox potentials of complexes 3 and the crystal structure of 3g (3,4-dichloro and N,N′-diisopropyl substituents) were determined.
Co-reporter:Tim Vorfalt Dipl.-Ing.;Klaus-Jürgen Wannowius Dr. Dr.
Angewandte Chemie International Edition 2010 Volume 49( Issue 32) pp:5533-5536
Publication Date(Web):
DOI:10.1002/anie.201000581
Co-reporter:Volodymyr Sashuk;LarsH. Peeck
Chemistry - A European Journal 2010 Volume 16( Issue 13) pp:3983-3993
Publication Date(Web):
DOI:10.1002/chem.200903275

Abstract

Imidazolium salts (NHCewgHCl) with electronically variable substituents in the 4,5-position (H,H or Cl,Cl or H,NO2 or CN,CN) and sterically variable substituents in the 1,3-position (Me,Me or Et,Et or iPr,iPr or Me,iPr) were synthesized and converted into the respective [AgI(NHC)ewg] complexes. The reactions of [(NHC)RuCl2(CHPh)(py)2] with the [AgI(NHCewg)] complexes provide the respective [(NHC)(NHCewg)RuCl2(CHPh)] complexes in excellent yields. The catalytic activity of such complexes in ring-closing metathesis (RCM) reactions leading to tetrasubstituted olefins was studied. To obtain quantitative substrate conversion, catalyst loadings of 0.2–0.5 mol % at 80 °C in toluene are sufficient. The complex with the best catalytic activity in such RCM reactions and the fastest initiation rate has an NHCewg group with 1,3-Me,iPr and 4,5-Cl,Cl substituents and can be synthesized in 95 % isolated yield from the ruthenium precursor. To learn which one of the two NHC ligands acts as the leaving group in olefin metathesis reactions two complexes, [(FL-NHC)(NHCewg)RuCl2(CHPh)] and [(FL-NHCewg)(NHC)RuCl2(CHPh)], with a dansyl fluorophore (FL)-tagged electron-rich NHC ligand (FL-NHC) and an electron-deficient NHC ligand (FL-NHCewg) were prepared. The fluorescence of the dansyl fluorophore is quenched as long as it is in close vicinity to ruthenium, but increases strongly upon dissociation of the respective fluorophore-tagged ligand. In this manner, it was shown for ring-opening metathesis ploymerization (ROMP) reactions at room temperature that the NHCewg ligand normally acts as the leaving group, whereas the other NHC ligand remains ligated to ruthenium.

Co-reporter:Dr. Tim Vorfalt;Dr. Klaus J. Wannowius;Dipl.-Chem. Vasco Thiel ;Dr. Herbert Plenio
Chemistry - A European Journal 2010 Volume 16( Issue 41) pp:12312-12315
Publication Date(Web):
DOI:10.1002/chem.201001832
Co-reporter:Jan Pschierer Dr.
Angewandte Chemie International Edition 2010 Volume 49( Issue 35) pp:6224-6227
Publication Date(Web):
DOI:10.1002/anie.201002045
Co-reporter:Lars H. Peeck, Steffen Leuthäusser, and Herbert Plenio
Organometallics 2010 Volume 29(Issue 19) pp:4339-4345
Publication Date(Web):August 25, 2010
DOI:10.1021/om100628f
Grubbs−Hoveyda and Grubbs III type complexes with ferrocenyl- or −NEt2-substituted NHC ligands were synthesized according to standard procedures. The electron donation of the NHC ligands in the respective ruthenium complexes can be modulated by oxidation of the ferrocenyl moiety or by protonation of the amino group. The neutral and the respective cationic (oxidized or protonated) ruthenium complexes were tested in the ROMP of norbornene. The change in the electron donation of the NHC ligands upon protonation leads to a significant change in the double-bond geometry (from E/Z ratio = 0.78 to E/Z = 1.04) and in the microstructure of the resulting polynorbornene. Consequently, addition of acid and protonation of the living catalyst attached to the polymer chain during the polymerization reaction allows fine-tuning the E/Z ratio of the resulting polynorbornene.
Co-reporter:Stefanie Wolf, Herbert Plenio
Journal of Organometallic Chemistry 2010 695(22) pp: 2418-2422
Publication Date(Web):
DOI:10.1016/j.jorganchem.2010.07.014
Co-reporter:Volodymyr Sashuk, Dirk Schoeps and Herbert Plenio  
Chemical Communications 2009 (Issue 7) pp:770-772
Publication Date(Web):14 Jan 2009
DOI:10.1039/B820633C
Fluorophore tagged N-heterocyclic carbenes and the derived (NHC)Pd(allyl)Cl complexes were synthesized and the fluorescence signal was used to follow the course of a Suzuki coupling reaction.
Co-reporter:Jan Pschierer and Herbert Plenio
Organic Letters 2009 Volume 11(Issue 12) pp:2551-2554
Publication Date(Web):May 18, 2009
DOI:10.1021/ol9007475
A general protocol is reported for the efficient Suzuki−Miyaura and the copper-free Sonogashira coupling of unprotected 6-chloropurines and unprotected β-d-ribofuranosyl-6-chloropurine in water or in water/n-butanol utilizing Na2PdCl4 and a disulfonated and highly water-soluble fluorenylphosphine (cataCXium F sulf).
Co-reporter:Dirk Schoeps, Volodymyr Sashuk, Katrin Ebert and Herbert Plenio
Organometallics 2009 Volume 28(Issue 13) pp:3922-3927
Publication Date(Web):June 15, 2009
DOI:10.1021/om900214j
An enlarged imidazolinium salt with a molecular mass of nearly 800 g/mol was synthesized and the respective N-heterocyclic carbene (NHC= N,N′-bis(2,6-diisopropyl-4-CH2NCy2-phenyl)-4,5-dihydroimidazol-2-ylidene) converted into (NHC)Pd(allyl)Cl and (NHC)Pd(cinnamyl)Cl complexes. The cinnamyl complex displays excellent activities in the Suzuki−Miyaura coupling and the Buchwald−Hartwig amination. The separation of this complex from the coupling products by means of a solvent-resistant nanofiltration using a PDMS (polydimethylsiloxane) membrane on PAN (polyacrylonitrile) was tested, and very high retention of between 97% and 99.9% of the (NHC)Pd complex was observed. The residual Pd content in the cross-coupling products is in the range 3.5−25 ppm.
Co-reporter:Stefanie Wolf, Herbert Plenio
Journal of Organometallic Chemistry 2009 694(9–10) pp: 1487-1492
Publication Date(Web):
DOI:10.1016/j.jorganchem.2008.12.047
Co-reporter:Tim Vorfalt Dipl.-Ing.;Steffen Leuthäußer Dr. Dr.
Angewandte Chemie International Edition 2009 Volume 48( Issue 28) pp:5191-5194
Publication Date(Web):
DOI:10.1002/anie.200900935
Co-reporter:Dirk Schoeps Dipl.-Ing.;Kristian Buhr Dipl.-Ing.;Marga Dijkstra Dr.;Katrin Ebert Dr. Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 12) pp:2960-2965
Publication Date(Web):
DOI:10.1002/chem.200802153
Co-reporter:Tim Vorfalt Dipl.-Ing.;Steffen Leuthäußer Dr. Dr.
Angewandte Chemie 2009 Volume 121( Issue 28) pp:5293-5296
Publication Date(Web):
DOI:10.1002/ange.200900935
Co-reporter:Christoph A. Fleckenstein and Herbert Plenio  
Green Chemistry 2008 vol. 10(Issue 5) pp:563-570
Publication Date(Web):11 Mar 2008
DOI:10.1039/B800154E
The water soluble Pd complex of dicyclohexyl(2-sulfo-9-(3-(4-sulfophenyl)propyl)-9H-fluoren-9-yl)phosphine was used for an efficient copper free and sustainable reaction protocol for Sonogashira cross couplings. Using a water/ipropanol mixture as the solvent and K2CO3 as base, numerous heterocyclic and aryl bromides and chlorides were reacted with various acetylenes in near quantitative yield at 1 mol% catalyst loading and 90 °C.
Co-reporter:Christoph A. Fleckenstein, Renat Kadyrov and Herbert Plenio
Organic Process Research & Development 2008 Volume 12(Issue 3) pp:475-479
Publication Date(Web):January 1, 2008
DOI:10.1021/op7001479
The reactions of aliphatic alcohols with fluorene coupled with a transfer hydrogenation result in the facile formation of 9-alkylfluorenes, whose deprotonation with nBuLi and quenching of the fluorenyl anion with Cy2PCl in MTBE gave 9-alkylfluorenyl-dicyclohexyl phosphines, which are conveniently isolated as the respective phosphonium tetrafluoroborates after treatment with aqueous HBF4. This route enables the facile large-scale (kilogram) synthesis of new ligands highly effective in Pd-catalyzed cross-coupling reactions.
Co-reporter:Herbert Plenio Dr.
Angewandte Chemie 2008 Volume 120( Issue 37) pp:7060-7063
Publication Date(Web):
DOI:10.1002/ange.200802270
Co-reporter:MarkusR. anderHeiden Dr. Dr.;Stefan Immel Dr.;Enrico Burello Dr.;Gadi Rothenberg Dr.;HuubC.J. Hoefsloot Dr.
Chemistry - A European Journal 2008 Volume 14( Issue 9) pp:2857-2866
Publication Date(Web):
DOI:10.1002/chem.200701418

Abstract

A method is presented for the high-throughput monitoring of reaction kinetics in homogeneous catalysis, running up to 25 coupling reactions in a single reaction vessel. This method is demonstrated and validated on the Sonogashira reaction, analyzing the kinetics for almost 500 coupling reactions. First, one-pot reactions of phenylacetylene with a set of 20 different meta- and para-substituted aryl bromides were analyzed in the presence of 17 different Pd–phosphine complexes. In addition, the temperature-dependent Sonogashira reactions were examined for 21 different ArX (X=Cl, Br, I) substrates, and the corresponding activation enthalpies and entropies were determined by means of Eyring plots: ArI (ΔH=48–62 kJ mol−1; ΔS=−71–−39 J mol−1 K; NO2OMe), ArBr (ΔH=54–82 kJ mol−1, ΔS=−55–11 J mol−1 K), and ArCl (ΔH=95–144 kJ mol−1, ΔS=−6–100 J mol−1 K). DFT calculations established a linear correlation of ΔH and the Kohn–Sham HOMO energies of ArX (X=Cl, Br, I) and confirmed their involvement in the rate-limiting step. However, despite different CX bond energies, aryl iodides and electron-deficient aryl bromides showed similar activation parameters.

Co-reporter:ChristophA. Fleckenstein Dipl.-Ing. Dr.
Chemistry - A European Journal 2008 Volume 14( Issue 14) pp:4267-4279
Publication Date(Web):
DOI:10.1002/chem.200701877

Abstract

A dicyclohexyl(2-sulfo-9-(3-(4-sulfophenyl)propyl)-9H-fluoren-9-yl)phosphonium salt was synthesized in 64 % overall yield in three steps from simple commercially available starting materials. The highly water-soluble catalyst obtained from the corresponding phosphine and [Na2PdCl4] enabled the Suzuki coupling of a broad variety of N- and S-heterocyclic substrates. Chloropyridines (-quinolines) and aryl chlorides were coupled with aryl-, pyridine- or indoleboronic acids in quantitative yields in water/n-butanol solvent mixtures in the presence of 0.005–0.05 mol % of Pd catalyst at 100 °C, chloropurines were quantitatively Suzuki coupled in the presence of 0.5 mol % of catalyst, and S-heterocyclic aryl chlorides and aryl- or 3-pyridylboronic acids required 0.01–0.05 mol % Pd catalyst for full conversion. The key to the high activity of the Pd-phosphine catalyst is the rational design of the reaction parameters (i.e., the presence of water in the reaction mixture, good solubility of reactants and catalyst in n-butanol/water (3:1), and the electron-rich and sterically demanding nature of the phosphine ligand).

Co-reporter:Herbert Plenio Dr.
Angewandte Chemie International Edition 2008 Volume 47( Issue 37) pp:6954-6956
Publication Date(Web):
DOI:10.1002/anie.200802270
Co-reporter:Holger Weychardt and Herbert Plenio
Organometallics 2008 Volume 27(Issue 7) pp:1479-1485
Publication Date(Web):March 8, 2008
DOI:10.1021/om701277p
1,4-Di(2-ethylhexyl)2,5-divinylbenzene, 1,4-di(2-ethylhexyloxy)-2,5-divinylbenzene, and 9,9-di(2-ethylhexyl)-2,7-divinylfluorene were synthesized from the respective dialdehydes and ADMET polymerized using Grubbs II or Grubbs−Hoveyda−Grela-type olefin metathesis catalysts. The continuous removal of ethene under a dynamic vacuum of 20 mbar at elevated temperatures in high boiling solvents (1,2-dichlorobenzene) resulted in the formation of long-chain poly[2,5-di(2-ethylhexyl)-p-phenylenevinylene] (Mn > 100 kDa, Pn = 330, Mw/Mn = 3.1) and poly[9,9-di(2-ethylhexyl)fluorenyl-2,7-vinylene] (Mn > 70 kDa, Pn = 178). The PPV forms a free-standing film after evaporation of a toluene solution. 1,3-(Diisopropenyl)ferrocene could not be polymerized in this manner; ferrocene-containing polymers were obtained with 1,1′-di(4-vinylphenyl)-3,3′,4,4′-tetramethylferrocene.
Co-reporter:Christoph A. Fleckenstein
Organometallics 2008 Volume 27(Issue 15) pp:3924-3932
Publication Date(Web):July 9, 2008
DOI:10.1021/om800259a
Seven new bidentate phosphine ligands, in which two 9-(dicyclohexylphosphino)fluoren-9′-yl units are bridged with n-alkanediyl- (C1−C5) or o/m-xylenediyl linkers were synthesized and characterized as the respective air-stable phosphonium salts. Pd complexes of the new bidentate ligands proved to be highly active catalysts for Buchwald−Hartwig amination and Suzuki and Sonogashira coupling using aryl bromides and chlorides as substrates. A study comparing the catalytic activity of the Pd complexes of the bidentate phosphine ligands with those of closely related monodentate fluorenyl phosphines gave insights into the influence of the various bridging units on the catalytic transformations. In Suzuki and amination reactions the diphosphine 2b with the shortest linker, a −CH2− unit, turned out to be by far the best ligand. In the Sonogashira coupling monodentate phosphines rendered the most active Pd catalysts, while 2b fails.
Co-reporter:Christoph A. Fleckenstein and Herbert Plenio  
Green Chemistry 2007 vol. 9(Issue 12) pp:1287-1291
Publication Date(Web):05 Oct 2007
DOI:10.1039/B711965H
The palladium complex of the new disulfonated 9-(3-phenylpropyl)-9′-PCy2-fluorene ligand is a highly active catalyst for aqueous Suzuki coupling reactions of N-heterocyclic chlorides and N-heterocyclic boronic acids; catalyst loadings of 0.02–0.1 mol% of Pd and two equiv. of phosphine result in the near quantitative formation of the respective coupling products at 100 °C.
Co-reporter:Christoph Fleckenstein, Sutapa Roy, Steffen Leuthäußer and Herbert Plenio  
Chemical Communications 2007 (Issue 27) pp:2870-2872
Publication Date(Web):02 May 2007
DOI:10.1039/B703658B
Sulfonated, water-soluble imidazolium and imidazolinium salts were synthesized and the respective Pd-complexes with N,N′-bis(2,6-dialkyl-4-SO3−-phenyl)imidazol-2-ylidene and N,N′-bis(2,6-dialkyl-4-SO3−-phenyl)-4,5-dihydroimidazol-2-ylidene ligands were applied in aqueous Suzuki coupling reactions of aryl chlorides.
Co-reporter:Markus an der Heiden and Herbert Plenio  
Chemical Communications 2007 (Issue 9) pp:972-974
Publication Date(Web):11 Dec 2006
DOI:10.1039/B616608C
The rates of Sonogashira coupling reactions using [Pd–PR3] complexes depend on a combination of the steric bulk of phosphines and substrates; however, below a critical cone angle of ca. 170° the catalytic activity drops drastically.
Co-reporter:Christoph A. Fleckenstein Dipl.-Ing. Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 9) pp:
Publication Date(Web):2 JAN 2007
DOI:10.1002/chem.200601142

The lithiation/alkylation of fluorene leads to various 9-alkyl-fluorenes (alkyl=Me, Et, iPr, -Pr, -C18H25) in >95 % yields, for which lithiation and reaction with R2PCl (R=Cy, iPr, tBu) generates 9-alkyl, 9-PR2-fluorenes which constitute electron-rich and bulky phosphine ligands. The in-situ-formed palladium–phosphine complexes ([Na2PdCl4], phosphonium salt, base, substrates) were tested in the Sonogashira, Suzuki, and Buchwald–Hartwig reactions of aryl chlorides and aryl bromides in organic solvents. The Sonogashira coupling of aryl chlorides at 100–120 °C leads to >90 % yields with 1 mol % of Pd catalyst. The Suzuki coupling of aryl chlorides typically requires 0.05 mol % of Pd catalyst at 100 °C in dioxane for quantitative product formation. To carry out “green” cross-coupling reactions in water, 9-ethylfluorenyldicyclohexylphosphine was reacted in sulphuric acid to generate the respective 2-sulfonated phosphonium salt. The Suzuki coupling of activated aryl chlorides by using this water-soluble catalyst requires only 0.01 mol % of Pd catalyst, while a wide range of aryl chlorides can be quantitatively converted into the respective coupling products by using 0.1–0.5 mol % of catalyst in pure water at 100 °C. Difficult substrate combinations, such as naphthylboronic acid or 3-pyridylboronic acid and aryl chlorides are coupled at 100 °C by using 0.1–0.5 mol % of catalyst in pure water to obtain the respective N-heterocycles in quantitative yields. The copper-free aqueous Sonogashira coupling of aryl bromides generates the respective tolane derivatives in >95 % yield.

Co-reporter:Steffen Leuthäußer Dipl.-Ing.;Daniela Schwarz  Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 25) pp:
Publication Date(Web):25 JUN 2007
DOI:10.1002/chem.200700228

The electron-donating properties of N-heterocyclic carbenes ([N,N′-bis(2,6-dimethylphenyl)imidazol]-2-ylidene and the respective dihydro ligands) with 4,4′-R-substituted aryl rings (4,4′-R=NEt2, OC12H25, Me, H, Br, S(4-tolyl), SO(4-tolyl), SO2(4-tolyl)) were studied. Twelve new N-heterocyclic carbene (NHC) ligands were synthesized as well as the respective iridium complexes [IrCl(cod)(NHC)] and [IrCl(CO)2(NHC)]. Cyclic voltammetry (ΔE1/2) and IR ((CO)) can be used to measure the electron-donating properties of the carbene ligands. Modifying the 4-positions with electron-withdrawing substituents (4-R=−SO2Ar, ΔE1/2=+0.92 V) results in NHC ligands with virtually the same electron-donating capacity as a trialkylphosphine in [IrCl(cod)(PCy3)] (ΔE1/2 =+0.95 V), while [IrCl(cod)(NHC)] complexes with 4-R=NEt2E1/2= +0.59 V) show drastically more cathodic redox potentials and significantly enhanced donating properties.

Co-reporter:Christoph A. Fleckenstein
Advanced Synthesis & Catalysis 2006 Volume 348(Issue 9) pp:
Publication Date(Web):28 JUN 2006
DOI:10.1002/adsc.200606024

The use of redox-switched phase tags in ferrocenyl-substituted triphenylphosphine combined with DBAD (di-tert-butyl azodicarboxylate) allows high yield (>90 %) Mitsunobu transformations without the need for the chromatographic purification of the products. The redox-switchable phosphine can be easily synthesized in two steps from 4-bromoaniline, ferrocene and chlorodiphenylphosphine. It is separated from the reaction mixture by oxidation with iron(III) chloride and can be recycled efficiently by reductive treatment.

Co-reporter:Marcus Süßner and Herbert Plenio  
Chemical Communications 2005 (Issue 43) pp:5417-5419
Publication Date(Web):30 Sep 2005
DOI:10.1039/B512008J
The donor properties of aryl substituted N-heterocyclic carbenes are characterized by lone pair donation from the carbene carbon and, as is shown here, by donation of electron density of the aromatic π-face of the NHC aryl groups towards the metal.
Co-reporter:Axel Köllhofer
Advanced Synthesis & Catalysis 2005 Volume 347(Issue 9) pp:
Publication Date(Web):19 JUL 2005
DOI:10.1002/adsc.200505095

A mixture of Na2PdCl4, CuI and (t-Bu)3PH+BF4 (molar ratio 4 : 3 : 8) dispersed in H2N(i-Pr)2+Br can be used as a “single source” precatalyst for the Sonogashira coupling of aryl bromides with various aryl- and alkylacetylenes in HN(i-Pr)2 solvent. Arylacetylenes require just 0.005 mol % of Pd catalyst at 80 °C, with TOFs ranging between 3,200 and 10,000 h−1.

Co-reporter:H. Plenio Dr.;A. Datta Dr.;M. an der Heiden Dipl.-Ing.;A. Köllhofer Dr.;S. Leuthäußer Dipl.-Ing.;M. Süßner Dipl.-Ing.
Chemie Ingenieur Technik 2005 Volume 77(Issue 8) pp:
Publication Date(Web):10 AUG 2005
DOI:10.1002/cite.200590258
Co-reporter:Marcus Süßner,Herbert Plenio
Angewandte Chemie International Edition 2005 44(42) pp:6885-6888
Publication Date(Web):
DOI:10.1002/anie.200502182
Co-reporter:Marcus Süßner Dipl.-Ing. Dr.
Angewandte Chemie 2005 Volume 117(Issue 42) pp:
Publication Date(Web):25 OCT 2005
DOI:10.1002/ange.200502182

Ein Homogenkatalysator wurde mit zwei Ferrocenylgruppen markiert, die sein Löslichkeitsverhalten durch reversibles Schalten zwischen neutralem und dikationischem Zustand bestimmen (siehe Schema). Mithilfe dieses Ansatzes wird die katalytische Aktivität eines Olefinmetathese-Katalysators effektiv an- und ausgeschaltet. Darüber hinaus sind die effiziente Trennung von Katalysator und Produkten sowie eine Wiederverwendung des Katalysators möglich.

Co-reporter:Anupama Datta, Axel Köllhofer and Herbert Plenio  
Chemical Communications 2004 (Issue 13) pp:1508-1509
Publication Date(Web):27 May 2004
DOI:10.1039/B403419H
The complex generated from ½ [Ir(OMe)(cod)]2 and 4,4′-di-tert-butyl-2,2′-bipyridine catalyzes the regioselective borylation of ferrocenes, CpMn(CO)3 and CpMo(CO)3CH3 with a stoichiometric amount of B2pin2.
Co-reporter:Herbert Plenio Dr.
ChemBioChem 2004 Volume 5(Issue 5) pp:
Publication Date(Web):28 APR 2004
DOI:10.1002/cbic.200300752

Fluorine can do it. Not only oxygen but also fluorine, covalently bonded in organic molecules, is an important donor atom in the coordination chemistry of biorelevant metal ions such as Na+, K+, Mg2+ and Ca2+. Clear evidence in favour of such interactions will be presented in this Minireview, followed by a discussion of possible implications for fluorinated, bioactive compounds.

Co-reporter:Markus an der Heiden Dipl.-Ing. Dr.
Chemistry - A European Journal 2004 Volume 10(Issue 7) pp:
Publication Date(Web):16 MAR 2004
DOI:10.1002/chem.200305562

The Suzuki–Miyaura coupling of aryl chlorides and PhB(OH)2 under biphasic conditions (DMSO/heptane) can be performed in almost quantitative yields over several cycles by means of polymeric Pd catalysts with soluble polyethylene glycol phase tags. Three sterically demanding and electron-rich phosphines 1-CH2Br,4-CH2P(1-Ad)2-C6H4, and 2-PCy2,2′-OH-biphenyl, and 2-PtBu2,2′-OH-biphenyl were covalently bonded to 2000 Dalton MeOPEG-OH. The catalysts, which were formed in situ from Na2[PdCl4], the respective polymeric phosphine, KF/K3PO4, and PhB(OH)2, efficiently couple aryl chlorides at 80 °C at 0.5 mol % loading, resulting in a >90 % yield of the respective biphenyl derivatives. The use of polar phase tags allows the efficient recovery of palladium–phosphine catalysts by simple phase separation of the catalyst-containing DMSO solution and the product-containing n-heptane phase. The high activity (TOF) of the catalyst remains almost constant over more than five reaction cycles, which involve the catalytic reaction, separation of the product phase from the catalyst phase, and addition of new reactants to initiate the next cycle. The Buchwald type biphenyl phosphines form the most active Pd catalysts, which are 1.3–2.8 times more active than catalysts derived from diadamantyl–benzylphosphine, but appear to be less robust in the recycling experiments. There is no apparent leaching of the catalyst into the heptane solution (<0.05 %), as evidenced by spectrophotometric measurements, and contamination of the product with Pd is avoided.

Co-reporter:Jens Hillerich and Herbert Plenio  
Chemical Communications 2003 (Issue 24) pp:3024-3025
Publication Date(Web):03 Nov 2003
DOI:10.1039/B310504K
The immobilisation of polymer tagged catalysts in a stationary solvent in a reactor and the flow of reactants dissolved in another immiscible solvent through it, allows the conversion of a continuous feed of reactants into a product stream, as exemplified by the Sonogashira coupling of aryl halides and acetylenes.
Co-reporter:Anupama Datta and Herbert Plenio  
Chemical Communications 2003 (Issue 13) pp:1504-1505
Publication Date(Web):21 May 2003
DOI:10.1039/B303602B
Palladium–phosphine catalysts which are phase-tagged with soluble poly(4-methylstyrene) can be used for efficient carbon–carbon coupling reactions by nonpolar biphasic catalysis with high recyclability as evidenced by high yields, constant tof of the catalyst and the absence of significant leaching of the catalyst into the polar product phase.
Co-reporter:Axel Köllhofer Dipl.-Chem. Dr
Chemistry - A European Journal 2003 Volume 9(Issue 6) pp:
Publication Date(Web):11 MAR 2003
DOI:10.1002/chem.200390161

The Sonogashira coupling of various aryl bromides and iodides with different acetylenes was studied under biphasic conditions with soluble, polymer-modified catalysts to allow the efficient recycling of the homogeneous catalyst. For this purpose, several sterically demanding and electron-rich phosphines of the type RPPR2 were synthesised. They are covalently linked to a monomethyl polyethylene glycol ether with a mass of 2000 Dalton (RP=MeOPEG2000) RPPR2: PR2=CH2C6H4CH2P(1-Ad)2, C6H4-P(1-Ad)2, C6H4-PPh2. To couple aryl iodides and acetylenes, the catalyst [(MeCN)2PdCl2]/2 RP-C6H4-PPh2 was used in CH3CN/Et3N/n-heptane (5/2/5). The combined yields of coupling product over five reaction cycles are between 80–95 %. There is no apparent leaching of the catalyst into n-heptane, as evidenced by 1H NMR spectroscopy. The new catalyst [(MeCN)2PdCl2]/2 (1-Ad)2PBn can be used for room-temperature coupling of various aryl bromides and acetylenes in THF with HNiPr2 as a base. A closely related catalyst Na2[PdCl4]/2 RP-CH2C6H4CH2P(1-Ad)2 linked to the polymer was used to couple aryl bromides and acetylenes in DMSO or DMSO/n-heptane at 60 °C with 0.5 mol % Na2[PdCl4], 1 mol % RPPR2 and 0.33 mol % CuI. The combined yield of coupling products over five cycles is always greater than 90 %, except for sterically hindered aryl bromides. The determination of the turnover frequency (TOF) of the catalyst indicates only a small decrease in activity over five cycles. Leaching of the catalyst into the product containing n-heptane solution could not be detected by means of 1H NMR and TXRF; this is indicative of >99.995 % catalyst retention in the DMSO solvent.

Co-reporter:Axel Köllhofer Dipl.-Chem.;Thomas Pullmann Dr.
Angewandte Chemie International Edition 2003 Volume 42(Issue 9) pp:
Publication Date(Web):26 FEB 2003
DOI:10.1002/anie.200390273

Aryl chlorides are suitable substrates for the Sonogashira coupling! By using the versatile catalyst system Na2[PdCl4]/PR3/CuI (PR3=(1-Ad)2PBn, PtBu3), the Sonogashira coupling [Eq. (a)] of aryl chlorides with alkynes generates excellent yields of the corresponding disubstituted aryl alkynes.

Co-reporter:Manuela Flauaus;Michael Herzing;Axel Köllhofer;Myriam Laly
European Journal of Organic Chemistry 2001 Volume 2001(Issue 6) pp:
Publication Date(Web):16 FEB 2001
DOI:10.1002/1099-0690(200103)2001:6<1061::AID-EJOC1061>3.0.CO;2-1

The (CH2Br)4 cavitand 1 and its reactions with NaOAc in varying stoichiometric ratios have been used to prepare a series of five different (CH2OAc)4−n(CH2Br)n cavitands (n = 0−3) 26 with mixed substituents, which can be separated by column chromatography (combined yield >95%) and are thus available in multi-gram quantities. The substitution reaction is statistical and the yields of individual acetates are controlled by the amount of NaOAc relative to 1. Careful hydrolysis of cavitands 26 with LiOH in THF-water results in the formation of the (CH2OH)4−n(CH2Br)n cavitands (n = 0−3) 711 in yields of between 70−85%, Treatment of 26 with thiourea and NaOH affords the respective (CH2SH)4−n(CH2OH)n cavitands (n = 0−3) 1216 in yields of around 90%. Reactions of 26 with K-Phthalimide generate the respective (CH2Pht.)4−n(CH2OAc)n cavitands (n = 0−3) 1721, which can be cleaved with hydrazine hydrate resulting in the (CH2NH2)4−n(CH2OH)n cavitands (n = 0−3) 2226 in an overall yield of 60−70%.

Co-reporter:Herbert Plenio Dr.;Clemens Aberle Dipl.-Chem.;Youseff Al Shihadeh;José Manuel Lloris Dr.;Ramón Martínez-Máñez Dr.;Teresa Pardo Dr.;Juan Soto Dr.
Chemistry - A European Journal 2001 Volume 7(Issue 13) pp:
Publication Date(Web):27 JUN 2001
DOI:10.1002/1521-3765(20010702)7:13<2848::AID-CHEM2848>3.0.CO;2-P

The reaction of 1,1′-ferrocene-bis(methylenepyridinium) salt with 1,4,8,11-tetraazacyclotetradecane-5,12-dione, followed by LiAlH4 reduction results in the formation of FcCyclam. Metal complexes of FcCyclam with M2+=Co2+, Ni2+, Cu2+, and Zn2+ were synthesized from FcCyclam and the respective metal triflates. The complexation of Cu2+ and FcCyclam in CH3CN is preceeded by a rapid electron transfer, followed by a slower complex formation reaction and a reverse electron transfer. The protonation constants of FcCyclam and the stability constants for the Cu2+ complex of FcCyclam (logK=9.26(4) for the formation of the [Cu(FcCyclam)]2+ complex) were determined in 1,4-dioxane/water 70:30 v/v, 0.1 mol dm−3, KNO3, 25 °C. By using FcCyclam one can selectively sense the presence of Cu2+ ions in the presence of Ni2+, Zn2+, Cd2+, Hg2+, and Pb2+ with a very large ΔE≈200 mV, depending on pH. The X-ray crystal structures of FcCyclam and of complexes with Co2+, Ni2+, Cu2+, and Zn2+ were determined and Fe−M2+ distances obtained: Fe−Co2+ 395.9, Fe−Ni2+ 385.4, Fe−Cu2+ 377.7, and Fe−Zn2+ 369.0 pm. The redox potential of FcCyclam is influenced in a characteristic manner by the complexation of M2+. A linear correlation of 1/r≅ΔE [r=distance Fe−M2+ from crystal data, ΔE=E1/2([M(FcCyclam)]2+)−E1/2(FcCyclam)] was found; this is indicative of a mainly Coulomb type interaction between the two metal centers. The nature of the Fe⋅⋅⋅M2+ interaction was also investigated by determining ΔE in several solvents (mixtures) of different dielectric constants ε. The expected relation of ΔE≅1/ε was only found at very high values of ε. At ε<40 increased ion-pairing appears to reduce the effective positive charge at M2+ leading to progessively smaller values of ΔE with lowered ε. The dependence of ΔE and ε can be calculated semiquantitatively by combining the Fuoss ion-pairing theory with the Coulomb model.

Co-reporter:Herbert Plenio Dr.;Clemens Aberle
Chemistry - A European Journal 2001 Volume 7(Issue 20) pp:
Publication Date(Web):5 OCT 2001
DOI:10.1002/1521-3765(20011015)7:20<4438::AID-CHEM4438>3.0.CO;2-K

1,1′-(OC2H4OTos)2-ferrocene was treated with various diaza-[n]-crown-m (n/m=12/4, 15/5, 18/6) to give three ferrocene cryptands (n/m=12/4 (FcCrypt), 15/5, 18/6). The complexation of Group 1 and 2 metal ions by FcCrypt leads to large shifts in the redox potentials (up to +500 mV relative to FcCrypt) and consequently to a drastic decrease in the binding strength (up to 108) in the ferrocene cryptands. The redox potential of Fcpda (1,1′-N,N′-bis(dipicol-2-ylamino)-3,3′,4,4′-tetraphenylferrocene) can be modified reversibly by complexation of Zn2+ (E(Fcdpa)=−0.13 V, E(Fcdpa-2 Zn+)=+0.66 V and E(Fcdpa-Zn2+)=+0.72 V). The X-ray crystal structure of FcCrypt-Ca(ClO4)2⋅H2O was determined; Ca2+ is coordinated by six oxygen (Ca2+−O 238.7, 239.1, 239.5, 242.6, 243.6, 247.7 pm) and two nitrogen donors (Ca2+−N 256.1, 259.2 pm) and displays a short Fe−Ca2+ contact (402.7 pm). The stability constants of FcCrypt-Na+ (lg K=8.32 in CH3CN) and FcCrypt-K+ (lg K=4.54 in CH3CN) were determined. The precise adjustment of complex stability and redox potentials of Fcdpa, Fcdpa-Zn2+, FcCrypt (+0.12 V), and FcCrypt-Na+ (+0.395 V) allows coupling of the redox-switchable ferrocene cryptand and the redox-responsive aminoferrocene. In a cyclic process starting from a mixture of Fcdpa+PF6 and FcCrypt-Na+ the addition of Zn(CF3SO3)2 raises the redox potential of Fcdpa+ to that of Fcdpa+-Zn2+. This complex oxidizes FcCrypt-Na+, while the oxidized cryptand displays a drastically reduced affinity towards Na+, so that a mixture containing FcCrypt+, Fcdpa-Zn2+, and uncoordinated Na+ results. The alkali metal ion is recomplexed after cyclam-assisted removal of Zn2+ from the Fcdpa-Zn2+ complex, since Fcdpa is oxidized by FcCrypt+ with reformation of FcCrypt-Na+. Thus two independent chemical processes—the complexation/decomplexation of Zn2+ and of Na+—are linked indirectly with mediation by electron-transfer reactions.

Co-reporter:Herbert Plenio Dr.;Jörg Hermann Dr.;Alexer Sehring
Chemistry - A European Journal 2000 Volume 6(Issue 10) pp:
Publication Date(Web):15 MAY 2000
DOI:10.1002/(SICI)1521-3765(20000515)6:10<1820::AID-CHEM1820>3.0.CO;2-A

The palladium-catalyzed Sonogashira reaction can be used to build optically active, oligomeric 1,2,3-substituted ferrocenes up to the tetramer, as well as polymers, by sequential coupling of optically active (ee>98 %), planar chiral iodoferroceneacetylenes and ferroceneacetylenes. (SFc)-1-Iodoferrocene-2-carbaldehyde (1) was reduced to the alcohol and methylated to give the corresponding methyl ether, which was Sonogashira-coupled with HC≡CSiEt3, resulting in (RFc)-1-(C≡CSiEt3)-2-methoxymethylferrocene (4) (79 %, three steps). Orthometalation with tBuLi followed by quenching with 1,2-diodoethane gave (RFc)-1-(C≡CSiEt3)-2-methoxymethyl-3-iodoferrocene (5). Deprotection of the acetylene with nBu4NF resulted in (RFc)-1-ethynyl-2-methoxymethyl-3-iodoferrocene (6), which was Sonogashira-coupled with itself to produce an optically active polymer. Deprotection of 4 with nBu4NF and Sonogashira coupling of the product with 5 resulted in the dinuclear ferrocene 9. Deprotection of 9 and coupling with 5, followed by deprotection of the resulting acetylene 11, gave the trinuclear ferrocene 12. Another such sequence involving 11 and 5 produced a tetranuclear ferrocene 13. To study the electronic communication in such oligomers in more detail, two symmetrical, closely interrelated, trinuclear ferrocenes 18 and 19 were synthesized. The redox potentials of all the ferrocenes and the ferroceneacetylene polymer were determined by cyclic and square-wave voltammetry. All the metallocenes were investigated by UV/Vis spectroscopy. A linear relationship was found between λmax and 1/n (n=number of ferrocene units in the oligomer). The polymer displayed two redox waves in the cyclic voltammogram, at 0.65 and 0.795 V. The corresponding mixed-valence oligoferrocene cations were synthesized from four ferroceneacetylenes, and their metal-metal charge transfer bands were examined by UV/Vis-NIR. The resonance exchange integrals Had, calculated on the basis of spectral information from the metal-metal charge transfer (MMCT) bands, were between 290 and 552 cm−1.

Co-reporter:Hans-Jürgen Buschmann;Jörg Hermann;Martin Kaupp
Chemistry - A European Journal 1999 Volume 5(Issue 9) pp:
Publication Date(Web):30 AUG 1999
DOI:10.1002/(SICI)1521-3765(19990903)5:9<2566::AID-CHEM2566>3.0.CO;2-3

The fluorine atom of fluorobenzene acts as a normal donor towards hard metal cations (see diagram). This conclusion was drawn on the basis of potentiometric and calorimetric data and X-ray crystal structure determinations. DFT calculations confirm this and predict the Li+–F (fluorobenzene) interaction to be roughly two-thirds as strong as that of Li+ with the oxygen atom of dimethyl ether.

Co-reporter:Volodymyr Sashuk, Dirk Schoeps and Herbert Plenio
Chemical Communications 2009(Issue 7) pp:NaN772-772
Publication Date(Web):2009/01/14
DOI:10.1039/B820633C
Fluorophore tagged N-heterocyclic carbenes and the derived (NHC)Pd(allyl)Cl complexes were synthesized and the fluorescence signal was used to follow the course of a Suzuki coupling reaction.
Co-reporter:Markus an der Heiden and Herbert Plenio
Chemical Communications 2007(Issue 9) pp:NaN974-974
Publication Date(Web):2006/12/11
DOI:10.1039/B616608C
The rates of Sonogashira coupling reactions using [Pd–PR3] complexes depend on a combination of the steric bulk of phosphines and substrates; however, below a critical cone angle of ca. 170° the catalytic activity drops drastically.
Co-reporter:Christoph Fleckenstein, Sutapa Roy, Steffen Leuthäußer and Herbert Plenio
Chemical Communications 2007(Issue 27) pp:NaN2872-2872
Publication Date(Web):2007/05/02
DOI:10.1039/B703658B
Sulfonated, water-soluble imidazolium and imidazolinium salts were synthesized and the respective Pd-complexes with N,N′-bis(2,6-dialkyl-4-SO3−-phenyl)imidazol-2-ylidene and N,N′-bis(2,6-dialkyl-4-SO3−-phenyl)-4,5-dihydroimidazol-2-ylidene ligands were applied in aqueous Suzuki coupling reactions of aryl chlorides.
Co-reporter:Christoph A. Fleckenstein and Herbert Plenio
Chemical Society Reviews 2010 - vol. 39(Issue 2) pp:NaN711-711
Publication Date(Web):2009/10/16
DOI:10.1039/B903646F
The strong electron-donation and the steric bulk of trialkylphosphines renders them as very useful ligands for palladium-catalyzed cross coupling reactions. This critical review reports on the synthesis of two families of trialkylphosphines (diadamantylalkylphosphines, fluorenyldialkylphosphines) and the properties of the respective palladium complexes in various cross coupling reactions, which evolved as alternatives to the classical Pd/PtBu3 system. In contrast to the latter phosphine the new classes of ligands are characterized by a highly flexible ligand design, which allows the fine tuning of catalytic properties to the specific needs of certain substrates and also enables the attachment of additional tags to impart certain useful properties onto the respective phosphines (179 references).
Co-reporter:Roman Savka, Sabine Foro and Herbert Plenio
Dalton Transactions 2016 - vol. 45(Issue 27) pp:NaN11024-11024
Publication Date(Web):2016/06/17
DOI:10.1039/C6DT01724J
The reaction of 1-amino,4-hydroxy-pentiptycene with diacetyl or acenaphthene-1,2-dione gave the respective diimines, followed by alkylation of the hydroxyl groups, and cyclization of the alkylated diimines to the respective bispentiptycene-imidazolium salts NHC·HCl. The azolium salts, being precursors to N-heterocyclic carbenes, were converted into metal complexes [(NHC)MX] (MX = CuI, AgCl, AuCl) and [(NHC)IrCl(cod)] and [(NHC)IrCl(CO)2] in good yields. In the solid state [(NHC)AgCl] displays a bowl-shaped structure of the ligand with the metal center buried within the concave unit.
Co-reporter:R. Savka and H. Plenio
Dalton Transactions 2015 - vol. 44(Issue 3) pp:NaN893-893
Publication Date(Web):2014/11/12
DOI:10.1039/C4DT03449J
The reactions of [MCl(cod)]2 (M = Rh, Ir) with different NHC·HX (X = Cl, I), K2CO3 in technical grade acetone under air provide simple access to various [(NHC)MX(cod)] complexes; a facile one-pot synthesis of [(NHC)MCl(CO)2] (M = Rh, Ir) is also reported.