Lutz H. Gade

Find an error

Name: Lutz H. Gade
Organization: Universit?t Heidelberg , Germany
Department: Anorganisch-Chemisches Institut
Title: (PhD)

TOPICS

Co-reporter:Christoph A. Rettenmeier, Hubert Wadepohl and Lutz H. Gade  
Chemical Science 2016 vol. 7(Issue 6) pp:3533-3542
Publication Date(Web):11 Feb 2016
DOI:10.1039/C5SC04644K
The study is aimed at a deeper understanding of the electronic structure of the T-shaped nickel(I) complex [LigiPr(iso)Ni] (1b), bearing the iso-PyrrMeBox (bis(oxazolinylmethylidene)pyrrolidinido) pincer ligand, and its CO adduct [LigiPr(iso)Ni(CO)] (2b) as well as to provide insight into the mechanism of autoxidation of the different nickel peroxo species of this ligand type. CO was found to react reversibly with complex 1b resulting in the corresponding CO adduct 2b. The EPR data as well as the results of DFT modeling revealed significant differences in the electronic structure of 1b and 2b. Reaction of [LigPh(iso)Ni] and [LigiPr(iso)Ni] (1a and b) with dioxygen yielded the 1,2-μ-peroxo complexes [Lig(iso)NiO]23a and b which reacted with hydrogen peroxide to give the hydroperoxo complexes [Lig(iso)NiOOH] 5a and b. Thermal aerobic decomposition of the peroxo species 3a and 5a in the presence of O2 led to a C–H activation of the ligand at the benzylic position of the oxazoline ring forming diastereomeric cyclic peroxo complexes 6 and 6′. For the 1,2-μ-peroxo complex 3b the autoxidation of the pincer in the absence of O2 occurred at the tertiary C–H bond of the iPr-group and led to a selective formation of the terminal hydroxo complex [LigiPr(iso)NiOH] 7b and the cyclic alkoxy complex 8 in equimolar quantities, while the corresponding cyclic peroxo species 9 was formed along with 7b in the presence of oxygen. Whether or not O–O bond cleavage occurred in the generation of 9 was established upon performing labeling experiments which indicate that the transformation does not involve an initial O–O bond cleaving step. Based on these observations and a series of stoichiometric transformations a tentative proposal for the processes involved in the anaerobic and aerobic decomposition of 3b has been put forward. Finally, the nickel(II) methyl complex [LigPh(iso)NiMe] 14 reacted with O2 to give the methylperoxo complex [LigPh(iso)NiOOMe] 15 which slowly converted to a mixture of near equal amounts of the formato and the hydroxo complexes, [LigPh(iso)NiOOCH] 16 and [LigPh(iso)NiOH] 7a, along with half an equivalent of methanol. The formato complex 16 itself decomposed at elevated temperatures to CO2, dihydrogen as well as the nickel(I) species 1a.
Co-reporter:Callum G. M. Benson, Alex J. Plajer, Raúl García-Rodríguez, Andrew D. Bond, Sanjay Singh, Lutz H. Gade and Dominic S. Wright  
Chemical Communications 2016 vol. 52(Issue 62) pp:9683-9686
Publication Date(Web):06 Jul 2016
DOI:10.1039/C6CC04805F
Deprotonation of the thialdiphosphazane [SPH(μ-NtBu)]2 with a range of metal-bases gives the stable dianion [S–P(μ-NtBu)]22−, which is valence-isoelectronic with the widely-used [RN-P(μ-NR)]22− ligand. Structural studies show that the new ligand has adaptable hard–soft character with respect to the coordinated metal centre and that its multidentate nature can be exploited to construct large cage architectures.
Co-reporter:Gudrun T. Plundrich, Hubert Wadepohl, and Lutz H. Gade
Inorganic Chemistry 2016 Volume 55(Issue 1) pp:353-365
Publication Date(Web):December 18, 2015
DOI:10.1021/acs.inorgchem.5b02498
This study focuses on the viability of the carbazole-based Cbzdiphos PNP pincer ligand as a stabilizing element for group 4 metal complexes, and both the diphenylphosphino- and di-isopropylphosphino-substituted Cbzdiphos protioligands 1PhH and 1iPrH were used. Treatment of the lithiated protioligands with the corresponding chlorido precursor compounds of the metals (titanium, zirconium, and hafnium) afforded the trichlorido complexes [(CbzdiphosiPr)MCl3] 2iPrM and [(CbzdiphosPh)MCl3] 2PhM (M = Ti, Zr, Hf), which were converted to the corresponding iodido complexes [(CbzdiphosiPr)MI3] 3iPrM and [(CbzdiphosPh)MI3] 3PhM (M = Ti, Zr, Hf) by reaction with an excess of trimethylsilyl iodide. Reaction of 2iPrTi and 3PhTi with 1 equiv of dibenzyl magnesium tetrahydrofuran adduct led to the formation of the alkylidene complexes 4iPrTi and 5PhTi, respectively, while the zirconium and hafnium complexes 2iPrZr and 3PhZr/Hf formed the cyclometalated monoalkyl compounds [(CbzdiphosiPr-CH)ZrBnCl] 6iPrZr as well as [(CbzdiphosPh-CH)MBnX] 6PhHf (X = Cl) and 7PhZr/Hf (X = I) under analogous reaction conditions. On the other hand, stirring 2PhZr with 0.25 equiv of tetrabenzyl zirconium afforded [(CbzdiphosPh)ZrBnCl2] (8PhZr), which contained the PNP ligand intact, while its alkylation with benzyl potassium led to the formation of the cyclometalated monobenzyl complex [(CbzdiphosPh-CH)ZrBnCl] (6PhZr). The remaining coordination site occupied by the halogenido ligand in the cyclometalated monobenzyl complexes [(Cbzdiphos-CH)MBnX] 6iPrZr, 6PhZr/Hf, and 7PhZr/Hf was readily benzylated by treatment with benzyl potassium to afford the cyclometalated dibenzyl complexes [(Cbzdiphos-CH)MBn2] 9iPrZr and 9PhZr/Hf. Further reaction of 9PhZr with an excess of benzyl potassium led to the formation of the anionic tribenzyl zirconium ate complex [(Cbzdiphos-CH)MBn3]K (10PhZr). Upon heating a solution of 8PhZr in the presence of 1 mol equiv of trimethyl phosphine, one of the ligand methylene groups was deprotonated, yielding the cyclometalated complex [(CbzdiphosPh-CH)ZrCl2(PMe3)] 11PhZr. Finally, reaction of 7PhZr with methylene triphenylphosphorane produced the ortho-metalated product [(CbzdiphosPh-CH)Zr(o-C6H4PPh2CH2)I] (12PhZr), which is characterized by a slightly puckered five-membered Zr–C(48)–P(3)–C(49)–C(50) metallacycle.
Co-reporter:Tim Bleith;Dr. Qing-Hai Deng; Hubert Wadepohl; Lutz H. Gade
Angewandte Chemie 2016 Volume 128( Issue 27) pp:7983-7987
Publication Date(Web):
DOI:10.1002/ange.201603072

Abstract

Whereas the stereochemical rigidity of the coordination sphere of boxmi/CuII catalysts is key to achieving high enantioselectivity in the electrophilic alkylation of β-ketoesters, this pathway is outperformed by a radical process for the corresponding catalytic transformation of oxindoles, giving rise to racemic products. For the corresponding ZnII catalysts, the selectivity in the latter process is outstanding despite the greater plasticity of the coordination shell. This reaction was thus developed into a highly useful synthetic method, which enabled the conversion of wide range of substrates with high yields and enantioselectivities.

Co-reporter:Tim Bleith;Dr. Qing-Hai Deng; Hubert Wadepohl; Lutz H. Gade
Angewandte Chemie International Edition 2016 Volume 55( Issue 27) pp:7852-7856
Publication Date(Web):
DOI:10.1002/anie.201603072

Abstract

Whereas the stereochemical rigidity of the coordination sphere of boxmi/CuII catalysts is key to achieving high enantioselectivity in the electrophilic alkylation of β-ketoesters, this pathway is outperformed by a radical process for the corresponding catalytic transformation of oxindoles, giving rise to racemic products. For the corresponding ZnII catalysts, the selectivity in the latter process is outstanding despite the greater plasticity of the coordination shell. This reaction was thus developed into a highly useful synthetic method, which enabled the conversion of wide range of substrates with high yields and enantioselectivities.

Co-reporter:Lena Hahn, Hubert Wadepohl, and Lutz H. Gade
Organic Letters 2015 Volume 17(Issue 9) pp:2266-2269
Publication Date(Web):April 22, 2015
DOI:10.1021/acs.orglett.5b00942
A new synthetic approach to core-functionalized tetraazaperopyrenes (TAPP) is reported. In-situ reaction of 4-fold lithiated TAPP with electrophiles results in the formation of various unprecedented TAPP derivatives, which are highly emissive fluorophores, show promising photophysical and electrochemical properties and act as valuable starting materials. Thus, lithiation of the TAPP core opens up a facile way for developing new organic materials.
Co-reporter:Astrid L. Müller, Hubert Wadepohl, and Lutz H. Gade
Organometallics 2015 Volume 34(Issue 12) pp:2810-2818
Publication Date(Web):March 20, 2015
DOI:10.1021/acs.organomet.5b00080
A series of osmium complexes with monoanionic, meridionally coordinating 1,3-bis(2-pyridylimino)isoindolates (BPI) as spectator ligands has been synthesized. Reaction of the dichlorido metal precursor [OsCl2(PPh3)3] with the lithiated BPI ligand transfer reagent gave the chlorido complex [(tBu-BPIMe)Os(PPh3)2Cl] (2) which, in turn, was reacted with lithium triethylborohydride to yield the hydrido complex [(tBu-BPIMe)Os(PPh3)2H] (3). Treatment of complex 2 with thallium hexafluorophosphate under a nitrogen pressure afforded the cationic dinitrogen complex [(tBu-BPIMe)Os(PPh3)2(N2)]PF6 (4), which contains an end-on coordinated dinitrogen molecule trans to the isoindolato nitrogen atom. To synthesize the alkynyl complex [(tBu-BPIMe)Os(PPh3)2(C≡CPh)] (5), the chlorido complex 2 was treated with lithium phenylacetylide. Complex 5 was subsequently converted quantitatively by addition of one or two equivalents of HBF4 to the vinylidene complex [(tBu-BPIMe)Os(PPh3)2(═C═CHPh)]BF4 (6), and the cationic species [(H-tBu-BPIMe)Os(PPh3)2(═C═CHPh)](BF4)2 (7), respectively; the latter being formed via protonation of one imine nitrogen atom of the BPI ligand. Upon stirring, a toluene solution of the chlorido complex 2 with benzyl potassium, the four-membered metallacycle [(tBu-BPIMe)Os(PPh3)(o-C6H4PPh2)] (8) was obtained, which reacted with molecular hydrogen and phenyl acetylene to give the hydrido complex 3 and the acetylide complex 5, respectively. Stirring of 8 with methyl acetylenedicarboxylate (DMAD) yielded [(tBu-BPIMe)Os(DMAD)(o-C6H4PPh2)] (9), while treatment with carbon monoxide gave the acyl complex [(tBu-BPIMe)Os(o–C(O)-C6H4PPh2)(CO)] (10) by an insertion of CO into the osmium carbon bond.
Co-reporter:Astrid L. Müller, Tim Bleith, Torsten Roth, Hubert Wadepohl, and Lutz H. Gade
Organometallics 2015 Volume 34(Issue 11) pp:2326-2342
Publication Date(Web):January 15, 2015
DOI:10.1021/om501138t
A series of κ2-(N,N)-coordinated bis(2-pyridylimino)isoindolato (BPI) complexes [Cp*Ir(BPI)Cl], which possess “three-legged piano-stool” structures, with the iridium atom being coordinated by the Cp* ligand 2 × N and Cl, were prepared via deprotonation of the BPIH ligands with either potassium hydride or LDA and subsequent reaction with [Cp*IrCl2]2 in THF. Cationic complexes [Cp*Ir(BPI)]+ containing κ3-(N,N,N)-coordinated BPI ligands were prepared as well as complexes with bidentate-coordinated BPI ligands, where the chloride ligand was substituted by either neutral or anionic ligands. Substitution in the ortho-position of the PBI ligands led to the formation of cycloiridated κ3-(N,N,C) species. Upon substitution of the anionic ligand by triphenylphosphine, a product was obtained with a hitherto unobserved κ2-(N,N) coordination of oMe-BPI to the metal center via the deprotonated nitrogen atom of the isoindole unit and one of the imine nitrogen atoms of the BPI ligand. A series of (para-cymene) osmium half-sandwich complexes with analogous structures and reactivities to their isoelectronic Cp*Ir(BPI) congeners were also prepared. Finally, it has been demonstrated that both Ir and Os complexes are catalytically active in the transfer hydrogenation of various ketones and imines.
Co-reporter:Christoph A. Rettenmeier; Hubert Wadepohl
Angewandte Chemie International Edition 2015 Volume 54( Issue 16) pp:4880-4884
Publication Date(Web):
DOI:10.1002/anie.201500141

Abstract

Pincer-stabilized nickel(I) complexes readily react with molecular oxygen to form dinuclear 1,2-μ-peroxo-bridged nickel(II) complexes, which are the major components of a dynamic equilibrium with the corresponding mononuclear superoxo species. The peroxo complexes further react with hydrogen peroxide to give the corresponding nickel(II) hydroperoxides. One of these hitherto elusive species was characterized by X-ray diffraction for the first time [O–O bond length: 1.492(2) Å].

Co-reporter:Yijing Xu, Christoph A. Rettenmeier, Gudrun T. Plundrich, Hubert Wadepohl, Markus Enders, and Lutz H. Gade
Organometallics 2015 Volume 34(Issue 20) pp:5113-5118
Publication Date(Web):October 2, 2015
DOI:10.1021/acs.organomet.5b00699
Reaction of the precursor complex [RuHCl(CO)(PPh3)3] with the PNP protioligand CbzdiphosH in toluene resulted in the formation of two stereoisomeric hydrido complexes, [(CbzdiphosH)RuHCl(CO)] (A). The addition of a strong base (KOtBu or LiEt3BH), on the other hand, led to the formation of the 1,2-dehydrochlorination product [(Cbzdiphos)RuH(CO)]. The reaction of the latter with BH3·THF at room temperature led to the 1,2-addition of the BH3 moiety to the Ru–N function, forming a RuNBH cycle in [(CbzdiphosHBH2)RuH(CO)] (B). The same borane-bridged compound was obtained when complex A was treated with NaBH4 in THF. The BH2 group forms a bridging unit between the carbazole-N atom and one of the ruthenium-bound hydrides.
Co-reporter:Christoph A. Rettenmeier; Hubert Wadepohl
Angewandte Chemie 2015 Volume 127( Issue 16) pp:4962-4966
Publication Date(Web):
DOI:10.1002/ange.201500141

Abstract

Pincer-stabilized nickel(I) complexes readily react with molecular oxygen to form dinuclear 1,2-μ-peroxo-bridged nickel(II) complexes, which are the major components of a dynamic equilibrium with the corresponding mononuclear superoxo species. The peroxo complexes further react with hydrogen peroxide to give the corresponding nickel(II) hydroperoxides. One of these hitherto elusive species was characterized by X-ray diffraction for the first time [O–O bond length: 1.492(2) Å].

Co-reporter:Susanne Langbein, Hubert Wadepohl, and Lutz H. Gade
The Journal of Organic Chemistry 2015 Volume 80(Issue 24) pp:12620-12626
Publication Date(Web):October 30, 2015
DOI:10.1021/acs.joc.5b01969
Fourfold arylmercapto substituted tetraazaperopyrene (TAPP) derivatives were obtained by direct nucleophilic substitution of the tetrabrominated TAPP or via Cu-catalyzed S–C coupling. These new materials display a characteristic bathochromic shift of their visible absorption and emission bands by ca. 200 nm compared to the unsubstituted parent compound. Two of the sulfide derivatives were oxidized with periodate to give their corresponding sulfones.
Co-reporter:Lena Hahn;Friedrich Maaß;Tim Bleith;Dr. Ute Zschieschang;Dr. Hubert Wadepohl;Dr. Hagen Klauk;Dr. Petra Tegeder;Dr. Lutz H. Gade
Chemistry - A European Journal 2015 Volume 21( Issue 49) pp:17691-17700
Publication Date(Web):
DOI:10.1002/chem.201503484

Abstract

A detailed study on the effects of core halogenation of tetraazaperopyrene (TAPP) derivatives is presented. Its impact on the solid structure, as well as the photophysical and electrochemical properties, has been probed by the means of X-ray crystallography, UV/Vis and fluorescence spectroscopy, high-resolution electron energy loss spectroscopy (HREELS), cyclic voltammetry (CV), and DFT modeling. The aim was to assess the potential of this approach as a construction principle for organic electron-conducting materials of the type studied in this work. Although halogenation leads to a stabilization of the LUMOs compared to the unsubstituted parent compound, the nature of the halide barely affects the LUMO energy while strongly influencing the HOMO energies. In terms of band-gap engineering, it was demonstrated that the HOMO–LUMO gap is decreased by substitution of the TAPP core with halides, the effect being found to be most pronounced for the iodinated derivative. The performance of the recently reported core-fluorinated and core-iodinated TAPP derivatives in organic thin-film transistors (TFTs) was investigated on both a glass substrate, as well as on a flexible plastic substrate (PEN). Field-effect mobilities of up to 0.17 cm2 Vs−1 and on/off current ratio of >106 were established.

Co-reporter:Qing-Hai Deng, Rebecca L. Melen, and Lutz H. Gade
Accounts of Chemical Research 2014 Volume 47(Issue 10) pp:3162
Publication Date(Web):August 28, 2014
DOI:10.1021/ar5002457
Tridentate monoanionic ligands known as “pincers” have gained a prominent place as ligands for transition metals and, more recently, for main-group metals and lanthanides. They have been widely employed as ancillary ligands for metal complexes studied inter alia in bond activation steps relevant to catalytic processes. The central formally anionic aryl or heteroaryl unit acts as an “anchor” in the coordination to the metal, which kinetically stabilizes the resulting complexes. Their stability, activity, and reactivity can be tuned by subtle modifications of substitution patterns on the pincer ligand or by modifying the donor atoms. The challenges in pincer ligand design for enantioselective catalysis have been met by their assembly from rigid heterocycles and chiral ligating units in the “wingtip” positions, which generally contain the stereochemical information. The resulting well-defined geometry and shape of the reactive sector of the molecular catalyst favor orientational control of the substrates. On the other hand, the kinetic stability allows reduced catalyst loadings.Recently, a new generation of tridentate anionic N∧N∧N pincer ligands has been developed which give rise to highly enantioselective transformations. Their applications in asymmetric catalysis have focused primarily on the asymmetric Nozaki–Hiyama–Kishi coupling of aldehydes with halogenated hydrocarbons as well as Lewis acid catalysis involving enantioselective electrophilic attack onto metal-activated β-keto esters, oxindoles, and related substrates. These include highly selective protocols for Friedel–Crafts alkylations with Michael acceptors, electrophilic fluorinations, trifluoromethylations, azidations, and alkylations and subsequent transformations. Increasingly, these stereodirecting ligands are being employed in other types of transformations, including hydrosilylations, cyclopropanations, and epoxidations. The stability and well-defined nature of the molecular catalysts have made them attractive targets for mechanistic studies into a wide range of these transformations, thus providing the type of insight required for a more rational approach to catalyst development. This Account reviews work performed by us and other groups in the field and places it into perspective in relation to general research efforts in enantioselective catalysis.
Co-reporter:Aneliia Shchyrba ; Christian Wäckerlin ; Jan Nowakowski ; Sylwia Nowakowska ; Jonas Björk ; Shadi Fatayer ; Jan Girovsky ; Thomas Nijs ; Susanne C. Martens ; Armin Kleibert ; Meike Stöhr ; Nirmalya Ballav ; Thomas A. Jung
Journal of the American Chemical Society 2014 Volume 136(Issue 26) pp:9355-9363
Publication Date(Web):June 24, 2014
DOI:10.1021/ja5020103
The formation of on-surface coordination polymers is controlled by the interplay of chemical reactivity and structure of the building blocks, as well as by the orientating role of the substrate registry. Beyond the predetermined patterns of structural assembly, the chemical reactivity of the reactants involved may provide alternative pathways in their aggregation. Organic molecules, which are transformed in a surface reaction, may be subsequently trapped via coordination of homo- or heterometal adatoms, which may also play a role in the molecular transformation. The amino-functionalized perylene derivative, 4,9-diaminoperylene quinone-3,10-diimine (DPDI), undergoes specific levels of dehydrogenation (−1 H2 or −3 H2) depending on the nature of the present adatoms (Fe, Co, Ni or Cu). In this way, the molecule is converted to an endo- or an exoligand, possessing a concave or convex arrangement of ligating atoms, which is decisive for the formation of either 1D or 2D coordination polymers.
Co-reporter:Lena Hahn, Simin Öz, Hubert Wadepohl and Lutz H. Gade  
Chemical Communications 2014 vol. 50(Issue 38) pp:4941-4943
Publication Date(Web):10 Mar 2014
DOI:10.1039/C4CC01254B
Water-soluble tetraazaperopyrene (TAPP) derivatives have been synthesized, which show both high photostability and high fluorescence quantum yields (>80%) in water. Furthermore delivery of the dye into cells demonstrated selective staining of the nucleus.
Co-reporter:Désirée C. Sauer;Rebecca L. Melen;Matthias Kruck
European Journal of Inorganic Chemistry 2014 Volume 2014( Issue 28) pp:
Publication Date(Web):
DOI:10.1002/ejic.201490140
Co-reporter:Désirée C. Sauer;Rebecca L. Melen;Matthias Kruck
European Journal of Inorganic Chemistry 2014 Volume 2014( Issue 28) pp:
Publication Date(Web):
DOI:10.1002/ejic.201402798

Abstract

Invited for the cover of this issue is the group of Professor L. H. Gade at the University of Heidelberg, Germany. The cover image shows the representation of a 1,3-bis(2-pyridylimino)isoindoline (BPI) compound along with its C2-symmetric derivatives. These ligands were first reported in the early 1950s and have found widespread application in organic, inorganic and materials chemistry.

Co-reporter:Désirée C. Sauer;Rebecca L. Melen;Matthias Kruck
European Journal of Inorganic Chemistry 2014 Volume 2014( Issue 28) pp:4715-N4725
Publication Date(Web):
DOI:10.1002/ejic.201402595

Abstract

Since the first report in the early 1950s, 1,3-bis(2-pyridylimino)isoindolines (BPIs) have found widespread applications in organic, inorganic and materials chemistry. This microreview focuses on recent progress towards chiral BPI derivatives as ligands for enantioselective catalysis as well as developments in the use of BPI complexes in materials science, focusing on luminescent and birefringent materials.

Co-reporter:Christoph Rettenmeier;Dr. Hubert Wadepohl ;Dr. Lutz H. Gade
Chemistry - A European Journal 2014 Volume 20( Issue 31) pp:9657-9665
Publication Date(Web):
DOI:10.1002/chem.201403243

Abstract

We herein report the catalytic enantioselective hydrodehalogenation based on the interplay of a chiral molecular nickel(I)/nickel(II)hydride system. Prochiral geminal dihalogenides are dehalogenated via a secondary configurationally unstable, potentially metal-stabilized radical intermediate. In a subsequent step, the liberated radical is then trapped by the nickel(II) hydrido complex, present in a large excess under the catalytic conditions, which in turn induces the enantioselectivity during the hydrogen atom transfer onto the radical intermediate. These new chiral nickel(I) complexes were found to catalyze the asymmetric hydrodehalogenation of geminal dihalogenides with moderate to good enantiomeric excess values using LiEt3BH as reductant. The main side product generally observed is the dehalogenated alkene, whereas the hydrodehalogenation of the chiral monohalogen compound occurred much more slowly despite the large excess of reductant.

Co-reporter:Dr. Qing-Hai Deng;Christoph Rettenmeier;Dr. Hubert Wadepohl;Dr. Lutz H. Gade
Chemistry - A European Journal 2014 Volume 20( Issue 1) pp:93-97
Publication Date(Web):
DOI:10.1002/chem.201303641

Abstract

The enantioselective trifluoromethylthiolation of β-ketoesters using chiral copper–boxmi complexes as catalysts is reported. A number of α-SCF3-substituted β-ketoesters have been obtained with up to >99 % enantiomeric excess (ee), and the trifluoromethylthiolated products were then transformed diastereoselectively to α-SCF3-β-hydroxyesters with two adjacent quaternary stereocenters.

Co-reporter:Qing-Hai Deng ; Tim Bleith ; Hubert Wadepohl
Journal of the American Chemical Society 2013 Volume 135(Issue 14) pp:5356-5359
Publication Date(Web):March 28, 2013
DOI:10.1021/ja402082p
The first example of Fe-catalyzed enantioselective azidations of β-keto esters and oxindoles using a readily available N3-transfer reagent is reported. A number of α-azido-β-keto esters were obtained with up to 93% ee, and this methodology also generates 3-substitued 3-azidooxindoles with high enantioselectivities (up to 94%).
Co-reporter:Sonja Geib;Ute Zschieschang;Marcel Gsänger;Matthias Stolte;Frank Würthner;Hubert Wadepohl;Hagen Klauk
Advanced Functional Materials 2013 Volume 23( Issue 31) pp:3866-3874
Publication Date(Web):
DOI:10.1002/adfm.201203600

Abstract

Organic thin-film transistors (TFTs) are prepared by vacuum deposition and solution shearing of 2,9-bis(perfluoroalkyl)-substituted tetraazaperopyrenes (TAPPs) with bromine substituents at the aromatic core. The TAPP derivatives are synthesized by reacting known unsubstituted TAPPs with bromine in fuming sulphuric acid, and their electrochemical properties are studied in detail by cyclic voltammetry and modelled with density functional theory (DFT) methods. Lowest unoccupied molecular orbital (LUMO) energies and electron affinities indicate that the core-brominated TAPPs should exhibit n-channel semiconducting properties. Current-voltage characteristics of the TFTs established electron mobilities of up to μn = 0.032 cm2 V−1 s−1 for a derivative which was subsequently processed in the fabrication of a complementary ring oscillator on a flexible plastic substrate (PEN).

Co-reporter:Solveig A. Scholl, Gudrun T. Plundrich, Hubert Wadepohl, and Lutz H. Gade
Inorganic Chemistry 2013 Volume 52(Issue 17) pp:10158-10166
Publication Date(Web):August 13, 2013
DOI:10.1021/ic401603y
The N-perfluoro-phenylated pyridyldiamine H2N2PFPNpy (1) has been prepared by a palladium-catalyzed coupling of hexafluorobenzene and the diamine (H2NCH2)2C(CH3)(2-C5H4N) using the palladacycle trans-di(μ-acetato)bis[o-(di-o-tolylphosphino)benzyl]palladium(II) as catalyst. Reactions of H2N2PFPNpy and Zr(NMe2)4 at room temperature or 90 °C led to the complexes [(NPFPN2TFAPNpy)ZrF(NMe2)] (2) and [(N2TFAPNpy)ZrF2] (3) in which one or two dimethylamido groups replaced one or two ortho fluorine atoms of the pentafluorophenyl groups in the ligand. Reaction of Me3SiX (X = Cl, I) with [(N2TFAPNpy)ZrF2] (3) resulted in the formation of mixed halogenated complexes [(N2TFAPNpy)ZrFI] (4) and [(N2TFAPNpy)ZrFCl] (5) in which the axially bound fluorido ligand is substituted. Reaction of [(N2TFAPNpy)ZrF2] (3) with LiNHNPh2 afforded the monohydrazido(1−) complex [(N2TFAPNpy)ZrF(NHNPh2)] (6) which was converted to the dimeric fluoro-potassium bridged hydrazinediido complex [Zr(N2TFAPNpy)FNNPh2K]2 (7) using KHMDS. The corresponding reaction with LiHMDS yielded the monomeric, donor free complex [Zr(N2TFAPNpy)NNPh2] (8).
Co-reporter:Nora Grüger, Lara-Isabel Rodríguez, Hubert Wadepohl, and Lutz H. Gade
Inorganic Chemistry 2013 Volume 52(Issue 4) pp:2050-2059
Publication Date(Web):January 30, 2013
DOI:10.1021/ic302454n
Two new monoanionic PNP pincer type ligands have been synthesized, the achiral 3,6-di-tert-butyl-1,8-bis((diphenyl-phosphino)methyl)-9H-carbazole CbzdiphosH (5) and the chiral 3,6-di-tert-butyl-1,8-bis(((2R,5R)-2,5-diphenylphospholan-1-yl)methyl)-9H-carbazole CbzdipholH (7), both of which were initially prepared as their borane complexes. The synthesis of CbzdiphosH is based on the reaction between the key intermediate 1,8-bis(bromomethyl)-3,6-di-tert-butyl-9H-carbazole (3) and lithium diphenylphosphide-borane complex. The chiral ligand CbzdipholH was prepared by treating 3 with lithium (2R,5R)-2,5-diphenylphospholanide-borane complex and subsequent deprotection with diethylamine. The complexation of the two ligands with nickel, palladium and rhodium was investigated, for which the conformational behavior of the ligands was found to be different. Although the arrangement of the donor atoms in all crystallographically characterized complexes is approximately square planar, the carbazole plane in Cbzdiphos complexes is inclined relative to the coordination plane. On the other hand, a helical twist is observed in Cbzdiphol complexes.
Co-reporter:Nora Grüger;Hubert Wadepohl
European Journal of Inorganic Chemistry 2013 Volume 2013( Issue 30) pp:5358-5365
Publication Date(Web):
DOI:10.1002/ejic.201300857

Abstract

The coordination chemistry of the monoanionic PNP pincer ligand 3,6-di-tert-butyl-1,8-bis[(diphenylphosphanyl)methyl]carbazole (Cbzdiphos) was investigated for different iridium(I) and iridium(III) complexes. Reaction of the protio ligand with [Ir(acac)(cod)] (acac = acetylacetonate; cod = 1,5-cyclooctadienyl) gave [(Cbzdiphos)Ir(cod)] (1). Upon hydrogenation the dihydrido complex [(Cbzdiphos)IrH2] (2) was obtained, and its reactivity in different organometallic transformations was examined. Treating 2 with ammonia afforded [(Cbzdiphos)IrH2(NH3)] (3), whereas treatment with diphenylacetylene (dpa) yielded the IrI compound [(Cbzdiphos)Ir(dpa)] (4) and trans-stilbene. Furthermore, the conversion with norbornene led to a reactive species, which activated C–H and C–Cl bonds. The reaction with THF gave a Fischer carbene [(Cbzdiphos)Ir{=C(OC3H6)}] (8) after double C–H bond activation, whereas treatment with benzene or chlorobenzene afforded the oxidative addition products [(Cbzdiphos)Ir(Ph)H] (5) and [(Cbzdiphos)Ir(Ph)Cl] (7).

Co-reporter:Achim Kruckenberg, Hubert Wadepohl, and Lutz H. Gade
Organometallics 2013 Volume 32(Issue 18) pp:5153-5170
Publication Date(Web):September 10, 2013
DOI:10.1021/om400711d
The facile one-step synthesis of five new bis(diisopropylphosphinomethyl)amine ligands RN(CH2DIP)2 (DIP = diisopropylphosphine, R = Me–, iPr–, PhCH2–, 2-ThCH2–, and 2-FuCH2−) based on the use of the air-stable phosphonium salt [DIP(CH2OH)2]Cl is presented. The phosphonium salt cleanly reacts with primary amines to afford amine-bridged bisphosphine ligands with variable backbone substitution in good yields. These DIP ligands are useful model systems for their chiral bisphospholane analogues. The coordination chemistry of neutral nickel(II) complexes [(RDIP)NiCl2], [(RDIP)Ni(CH3)2], [(iPrDIP)Ni(Cl)CH2Si(CH3)3], and [(RDIP)Ni(CH2Ph)2], as well as coordinatively unsaturated cationic nickel(II) complexes [(RDIP)Ni(THF)CH3]+BArF– and [(RDIP)NiCH2Ph]+BArF–, has been studied by spectroscopic and X-ray diffraction methods. The cationic methyl complexes reacted cleanly with 2-butyne and 1,1-dimethylallene, yielding allylic complexes [(RDIP)Ni(pmcb)]+BArF– and [(RDIP)Ni(tma)]+BArF–, respectively (pmcb: η3-1,2,3,4,4-pentamethylcyclobutenyl; tma: η3-2,3,3-trimethylallyl). Nickel(0) complexes [(BzDIP)Ni(trans-stilbene)] and [(RDIP)Ni(C2H4)] were synthesized in one step from the corresponding dichloro complexes and have been fully characterized and analyzed by X-ray diffraction methods. [(MeDIP)Ni(C2H4)] reacted with phenyl-2-thiophenecarboxylate, yielding [(MeDIP)Ni(OPh)(2-Th)] and [(MeDIP)Ni(CO)2] in a 2:1 ratio. Both complexes were also synthesized via alternative routes. The phenolato-thienyl complex represents an intermediate in the catalytic cycle of a recently reported decarbonylative arylation of azoles. Complexes [(RDIP)Ni(C2H4)] were shown to be active catalysts for this reaction.
Co-reporter:Peter Scherl, Hubert Wadepohl, and Lutz H. Gade
Organometallics 2013 Volume 32(Issue 15) pp:4409-4415
Publication Date(Web):July 22, 2013
DOI:10.1021/om400574a
A double-cyclometalated ruthenium complex containing a chiral tripodal phospholane has been prepared by reaction with [Ru(η4-COD)(η3-methylallyl)2] via elimination of isobutene. The ruthenium–carbon bonds of this compound were reversibly cleaved by H2, resulting in an equilibrium between a tri- and a tetrahydride (4 and 5). T1 relaxation time measurements revealed the nonclassical nature of the fluctuating hydrides. Release of the gas led to complete re-formation of the cyclometalated compound. Reaction of 3 with D2 afforded D10-5, in which six ortho-phenyl protons and four hydrides were replaced by deuterium. Furthermore, diphenylsilane was found to readily insert into one Ru–C bond to form 6, containing a κ3-dihydridosilicate fragment. On the basis of deuterium labeling experiments, the fast exchange between the two hydrides was shown to include a reductive elimination/oxidative addition step involving the remaining metalated phenyl group. Again, pressurization of 6 with H2 resulted in reversible cleavage of the remaining Ru–C bond, yielding the corresponding trihydride 7.
Co-reporter:Thorsten Gehrmann, Julio Lloret-Fillol, Heike Herrmann, Hubert Wadepohl, and Lutz H. Gade
Organometallics 2013 Volume 32(Issue 14) pp:3877-3889
Publication Date(Web):July 8, 2013
DOI:10.1021/om400337g
The N–N bond in the zirconium hydrazinediido(2−) complex [Zr(N2TBSNpy)(NNPh2)(py)] (1) is readily cleaved by one-electron oxidation. Reacting [Zr(N2TBSNpy)(NNPh2)(py)] (1) with 0.5 molar equiv of iodine led to the release of molecular N2 and yielded the mixed diphenylamido/iodo complex [Zr(N2TBSNpyNPh2)(I)] (2). Exposure of hydrazinediide 1 to an excess of iodine resulted in further oxidation of the diphenylamido ligand, yielding the diiodo complex 3 and tetraphenylhydrazine. Similar reactivity was observed in the reaction of 1 with diphenyl diselenide and diaryl disulfides, which reacted to give the corresponding diphenylamido/arylchalcogenido complexes [Zr(N2TBSNpyNPh2)(SePh)] (4a) and [Zr(N2TBSNpy)(NPh2)(SAr)] (Ar = Ph (4b), C6F5 (4c)) along with N2. The reactions were also carried out on an NMR scale with a 15Nα-labeled hydrazido complex (1-15N). In all cases a single 15N NMR resonance at 310.16 ppm, assigned to 15N2, indicated the formation of dinitrogen from the Nα atom in the hydrazide. A crossover labeling experiment employing a 1:1 mixture of 1 and 15Nα-labeled 1-15N revealed that the isotope distribution is, as expected, statistical 1:2:1 (14N2: 14/15N2: 15N2), which is consistent with a reaction pathway involving a dinuclear intermediate in the dinitrogen-forming step. Complex 1 reacted with N2O to give a mixture of two compounds, the bis(diphenylamido) complex 6 and the doubly bridged μ-oxo complex 7. In contrast, reaction of 1 with 1 molar equiv of pyridinium N-oxide only gave the doubly bridged μ-oxo complex 7 along with 2,2′-bipyridine and diphenylamine.
Co-reporter:Désirée C. Sauer, Matthias Kruck, Hubert Wadepohl, Markus Enders, and Lutz H. Gade
Organometallics 2013 Volume 32(Issue 3) pp:885-892
Publication Date(Web):January 22, 2013
DOI:10.1021/om301198b
Iron(II) and cobalt(II) alkyl complexes using tridentate bis(pyridylimino)isoindolates as ancillary ligands have been synthesized from the pyridine alkyl precursor complexes [(py)2Fe(CH2SiMe3)2] and [(py)2Co(CH2SiMe3)2]. The extremely air- and moisture-sensitive compounds were structurally characterized in the solid state by X-ray diffraction as well as in solution by paramagnetic NMR spectroscopy. It is demonstrated that the paramagnetic shifts in the 13C NMR spectra are dominated by strong Fermi-contact interactions. All 13C NMR signals can be assigned by correlation with DFT-calculated spin-density distributions.
Co-reporter:Solveig A. Scholl, Hubert Wadepohl, and Lutz H. Gade
Organometallics 2013 Volume 32(Issue 3) pp:937-940
Publication Date(Web):January 27, 2013
DOI:10.1021/om301235d
Reaction of cyclic 1,1′-disubstituted hydrazines with the bis(dimethylamido)zirconium complex [Zr{(NXyl)2Npy} (NMe2)2] (1) in the presence of dmap yielded the hexacoordinate zirconium hydrazinediido complexes [Zr{(NXyl)2Npy}(═NNC9H10)(dmap)2] (2) and [Zr{(NXyl)2Npy}(═NNC12H8)(dmap)2] (3). Hydrazinediides are thought to be key intermediates in the zirconium-catalyzed reaction of cyclic 1,1′-disubstituted hydrazines and disubstituted alkynes to yield 1,7-annulated indoles. Their basic structural motif is found in 5-HT3 receptor antagonists such as Cilansetron.
Co-reporter:Matthias Kruck;Dr. Hubert Wadepohl;Dr. Markus Enders;Dr. Lutz H. Gade
Chemistry - A European Journal 2013 Volume 19( Issue 5) pp:
Publication Date(Web):
DOI:10.1002/chem.201390010
Co-reporter:Matthias Kruck;Dr. Hubert Wadepohl;Dr. Markus Enders;Dr. Lutz H. Gade
Chemistry - A European Journal 2013 Volume 19( Issue 5) pp:1599-1606
Publication Date(Web):
DOI:10.1002/chem.201203450

Abstract

High-spin FeII–alkyl complexes with bis(pyridylimino)isoindolato ligands were synthesized and their paramagnetic 1H and 13C NMR spectra were analyzed comprehensively. The experimental 13C—1H coupling values are temperature (T−1)- as well as magnetic-field (B2)-dependent and deviate considerably from typical scalar 1JCH couplings constants. This deviation is attributed to residual dipolar couplings (RDCs), which arise from partial alignment of the complexes in the presence of a strong magnetic field. The analysis of the experimental RDCs allows an unambiguous assignment of all 13C NMR resonances and, additionally, a structural refinement of the conformation of the complexes in solution. Moreover the RDCs can be used for the analysis of the alignment tensor and hence the tensor of the anisotropy of the magnetic susceptibility.

Co-reporter:Dr. Sonja Geib;Dr. Susanne C. Martens;Michaela Märken;Arina Rybina; Hubert Wadepohl; Lutz H. Gade
Chemistry - A European Journal 2013 Volume 19( Issue 41) pp:13811-13822
Publication Date(Web):
DOI:10.1002/chem.201301903

Abstract

Core substitution of tetraazaperopyrenes (TAPPs) has been achieved, and with it, considerable variation of their photo- and redox-chemical properties. Through Suzuki cross coupling starting from the fourfold core-brominated tetraazaperopyrene, aryl-substituted TAPPs were synthesized, which displayed very high fluorescence quantum yields (up to 100 %) in solution. Besides the Suzuki reactions, Stille and Sonogashira cross-couplings were also found to be suitable methods for core derivatization, as demonstrated in the syntheses of alkynyl-substituted tetraazaperopyrene congeners. Furthermore, TAPPs incorporating intramolecular donor–acceptor combinations of aromatic units (8, 9) were accessible by coupling the electron-poor peropyrene core with electron-rich aromatic units, which act as strong electron donors. Finally, C-heteroatom coupling (O, S, N) gave rise to novel TAPP derivatives with strongly modified redox-chemical behaviour and photophysics in the solid state as well in solution. In particular, TAPP derivatives displaying red fluorescence were obtained for the first time.

Co-reporter:Dipl.-Chem. Torsten Roth;Dr. Hubert Wadepohl;Dr. Dominic S. Wright;Dr. Lutz H. Gade
Chemistry - A European Journal 2013 Volume 19( Issue 41) pp:13823-13837
Publication Date(Web):
DOI:10.1002/chem.201302327

Abstract

A series of dichlorocyclophosphazanes [{ClP(μ-NR)}2] containing chiral and achiral R groups was obtained from simple commercially available amines and PCl3. Their condensation reactions with axially chiral biaryl diols yielded ansa-bridged chiral cyclophosphazane (CycloP) ligands. This highly modular methodology allows extensive elaboration of the ligand set, in which the chirality can be introduced at the diol bridge and/or the amido R group. This provides the possibility to observe match and mismatch effects in catalysis. A series of twenty CycloP ligands was synthesized and characterized by multinuclear NMR spectroscopy, HRMS, elemental analysis, and in selected cases, single-crystal X-ray diffraction. These studies show that all of the ditopic CycloP ligands are C2 symmetric, rendering their metal coordination sites symmetry equivalent. Two well-established enantioselective reactions were explored by using late-transition metal CycloP complexes as catalysts; the gold-catalyzed hydroamination of γ-allenyl sulfonamides and the asymmetric nickel-catalyzed three-component coupling of a diene and an aldehyde. The steric demands of the CycloP ligands have a subtle influence on the reactivity and selectivity observed in both reactions. Good enantiomeric ratios (e.r.) as high as 89:11 in the gold-catalyzed reaction and 92:8 in the nickel-catalyzed bis-homoallylation reaction were observed.

Co-reporter:Qing-Hai Deng ; Hubert Wadepohl
Journal of the American Chemical Society 2012 Volume 134(Issue 6) pp:2946-2949
Publication Date(Web):January 26, 2012
DOI:10.1021/ja211859w
Cu-catalyzed enantioselective alkylation of β-ketoesters using alcohols for in situ preparation of alkylating reagents is reported. A number of functionalized β-ketoesters containing a quaternary carbon stereocenter are obtained with up to 99% ee. The alkylation products derived from 2-substituted allylic alcohols or their corresponding iodides can then be converted to spirolactones, bi-spirolactones, and related chiral target products.
Co-reporter:Qing-Hai Deng ; Hubert Wadepohl
Journal of the American Chemical Society 2012 Volume 134(Issue 26) pp:10769-10772
Publication Date(Web):June 13, 2012
DOI:10.1021/ja3039773
Enantioselective Cu-catalyzed trifluoromethylation of β-ketoesters using commercially available trifluoromethylating reagents is reported. A number of α-CF3 β-ketoesters are obtained with up to 99% ee. The trifluoromethylated products were then transformed diastereoselectively to α-CF3 β-hydroxyesters with two adjacent quaternary stereocenters via a Grignard reaction.
Co-reporter:Thorsten Gehrmann, Matthias Kruck, Hubert Wadepohl and Lutz H. Gade  
Chemical Communications 2012 vol. 48(Issue 18) pp:2397-2399
Publication Date(Web):25 Jan 2012
DOI:10.1039/C2CC17594K
Reaction of a zirconium imido and a hydrazinediido complex with bis(trimethylsilyl)sulfur diimide yielded the corresponding formal [2+2] cycloaddition products, the trisimidosulfito complex and the hydrazidobis(imido)sulfito complex containing an unprecedented SN2(N–N) unit.
Co-reporter:Désirée C. Sauer, Hubert Wadepohl, and Lutz H. Gade
Inorganic Chemistry 2012 Volume 51(Issue 23) pp:12948-12958
Publication Date(Web):November 15, 2012
DOI:10.1021/ic3020749
The synthesis of a new family of chiral tridentate monoanionic NNN-pincer ligands based on the 1,3-bis(2-pyridylimino)isoindoline (BPI) framework is reported. Ligands with substituents of varying steric demand were prepared starting from achiral and low priced materials. A kinetic enzymatic resolution was used as a key step for the preparation of enantiomerically pure ligands. In this way, both enantiomers of a given ligand could be produced enantioselectively (>99.5% ee). The corresponding cobalt alkyl complexes were obtained using a pyridine alkyl cobalt precursor complex and were applied in asymmetric hydrosilylation of several prochiral alkylaryl ketones with high yields (up to 100%) and enantioselectivity (up to 91% ee) to give the chiral alcohols after hydrolysis.
Co-reporter:Nora Grüger, Hubert Wadepohl and Lutz H. Gade  
Dalton Transactions 2012 vol. 41(Issue 46) pp:14028-14030
Publication Date(Web):12 Oct 2012
DOI:10.1039/C2DT32199H
The novel chelating PNP pincer ligand 2,5-bis((diphenylphosphino)methyl)-1H-pyrrole (1) was prepared and its nickel coordination chemistry explored. The reaction with Ni(COD)2 led to a diamagnetic dinuclear nickel(I) complex (4) which was also obtained by the reaction of the square planar NiII complexes [(PNP-Ph2)NiX] (X = Cl (2), X = I (3)) with Li(Et3BH).
Co-reporter:Yan-Biao Kang and Lutz H. Gade
The Journal of Organic Chemistry 2012 Volume 77(Issue 3) pp:1610-1615
Publication Date(Web):January 9, 2012
DOI:10.1021/jo202491y
A clean and efficient diacetoxylation reaction of alkenes catalyzed by triflic acid using commercially available peroxyacids as the oxidants has been developed. This method was also applied in oxidative lactonizations of unsaturated carboxylic acids in good to high yields.
Co-reporter:Peter Scherl, Achim Kruckenberg, Steffen Mader, Hubert Wadepohl, and Lutz H. Gade
Organometallics 2012 Volume 31(Issue 19) pp:7024-7027
Publication Date(Web):September 20, 2012
DOI:10.1021/om300794h
Ruthenium η4-trimethylenemethane complexes containing two different tridentate phosphine ligands have been synthesized. The formation of the complexes using [Ru(η4-COD)(η3-CH2CMeCH2)2] as the metal precursor occurred via elimination of isobutene. An intermediate species has been isolated in which the ligand coordinates with only two phosphorus atoms. Furthermore, protonation of the trimethylenemethane ligand gave rise to a cationic methylallyl complex.
Co-reporter:Thorsten Gehrmann, Julio Lloret Fillol, Hubert Wadepohl, and Lutz H. Gade
Organometallics 2012 Volume 31(Issue 12) pp:4504-4515
Publication Date(Web):June 4, 2012
DOI:10.1021/om300291b
Reaction of the zirconium dichloro complex [Zr(N2TBSNpy)Cl2] (1) with 1 molar equiv of ArNHLi (Ar = Mes, DIPP) yielded the zirconium imido complexes [Zr(N2TBSNpy)(═NDIPP)(py)] (2; N2TBSNpy = [(2-C5H4N)C(CH3){CH2NSi(CH3)2tBu}2]2–, DIPP = 2,6-diisopropylphenyl) and [Zr(N2TBSNpy)(═NMes)(py)] (3; Mes = mesityl). The imido complexes are converted to the tetraazadienido complexes [Zr(N2TBSNpy)(NDIPPN2NPh)] (4) and Zr(N2TBSNpy)(NMesN2NPh)] (5) by addition of phenyl azide, whereas the reaction of 2 or 3 with mesityl azide gave the alternative product 7, in which the azide is coupled with the CH activated ancillary tripod ligand. Reaction of 1 molar equiv of trimethylsilyl azide or 1-adamantyl azide with the previously reported hydrazinediido complex [Zr(N2TBSNpy)(═NNPh2)(py)] (9) at ambient temperature resulted in the formation of the five-membered zirconaacacycles [Zr(N2TBSNpy)(NTMSN3NPh2)] (10) and [Zr(N2TBSNpy)(NAdN3NPh2)] (11). Complex 11 was thermally converted into the diazenido complex 12 via loss of 1 molar equiv of molecular N2. The direct formation of the analogous side-on-bonded diazenido analogue 13 was observed upon reaction of 9 with 1 equiv of mesityl azide at ambient temperature. On the basis of 15N labeling and DFT modeling (DFT(B3PW91/6-31 g(d))) a mechanism for the reaction pathway leading to 12 and 13 is proposed.
Co-reporter:Thorsten Gehrmann, Gudrun T. Plundrich, Hubert Wadepohl, and Lutz H. Gade
Organometallics 2012 Volume 31(Issue 8) pp:3346-3354
Publication Date(Web):April 3, 2012
DOI:10.1021/om300143h
Reactions of the hydrazinediido complexes [M(N2TBSNpy)(NNPh2)(py)] (M = Zr (1a), Hf (1b)) with (hetero)allenes result in a variety of [2 + 2] cycloaddition products of the general type [M(N2TBSNpy)(κ2N,E-(E(═E′R)NNPh2)(py)] (E = CH2, S; E′ = CH, N; R = alkyl, aryl). The reaction of [Zr(N2TBSNpy)(NNPh2)(py)] (1a) with 1 molar equiv of phenyl or mesityl isothiocyanate at room temperature yields [Zr(N2TBSNpy)(κ2N,S-SC(═NAr)NNPh2)(py)] (Ar = phenyl (2a), mesityl (2b)). Reacting the hydrazinediides [M(N2TBSNpy)(NNPh2)(py)] (M = Zr (1a), Hf (1b)) with allenes results in the formation of the metallaazacyclobutanes [M(N2TBSNpy)(κ2N,C-N(NPh2)CH2C═CH(R))(py)] (M = Zr, R = Ph (4a), cyclohexyl (5a), methyl (6); M = Hf, R = phenyl (4b), cyclohexyl (5b)). Subsequent heating of the cycloaddition products revealed different reactivity patterns: the complex [Zr(N2TBSNpy)(κ2N,S-SC(═NAr)NNPh2)(py)] (2a) forms the isomerization product [Zr(N2TBSNpy)(κ2N,S-SC(═NNPh2))NPh] (3), retaining the N–N bond of the hydrazide. In contrast, the metallacyclobutanes 4a,b and 5a,b show a tendency toward N–N bond cleavage, resulting in the formation of the C–N- and C–C-coupled product complexes [M(κ4N,N,N,N-N2TBSNpyNC(Me)═CHCy)(NPh2)] (M = Zr (7a), Hf (7b)), [Zr(N2TBSNpy)(κ2N,C-(Ph)NC6H4C(Me)═C(Ph)NH)] (8) and [Zr(κ4N,N,N,N-N2TBSNpyNC(Me)=CHPh)(NPh2)] (9).
Co-reporter:Dr. Thorsten Gehrmann;Solveig A. Scholl;Dr. Julio Lloret Fillol;Dr. Hubert Wadepohl ;Dr. Lutz H. Gade
Chemistry - A European Journal 2012 Volume 18( Issue 13) pp:3925-3941
Publication Date(Web):
DOI:10.1002/chem.201103497

Abstract

Reaction of [Zr{(NAr)2Npy}(NMe2)2] (Ar=3,5-xylyl: 2 a, mesityl: 2 b) with one or two molar equivalents of 1,1-diphenylhydrazine gave the mixed amido/hydrazido(1−) complex [Zr{(NMes)2Npy}(HNNPh2)(NMe2)] (3), the bis-hydrazido complex [Zr{(NMes)2Npy}(HNNPh2)2] (4), and, in the presence of excess 4-dimethylaminopyridine (DMAP), hexacoordinate hydrazinediidozirconium complexes [Zr{(NXyl)2Npy}(NN(Me)Ph)(dmap)2] (5) and [Zr{(NXyl)2Npy}(NNPh2)(dmap)2] (6). The reaction of one equivalent of the zirconium–hydrazinediide [Zr{(NTBS)2Npy}(NNPh2)(py)] (1) with disubstituted alkynes at RT for 16 h led to the formation of seven-membered diazazirconacycles 7 a7 e in high yields. Similar reactivity was observed by reacting bis-amido complex 2 b with one molar equivalent of the corresponding alkyne and diphenylhydrazine. The formation of the seven-membered zirconacycles implied a key coupling step that involved the alkyne and one of the aryl rings of the diphenylhydrazinediido ligand. In some cases, such as the reaction with 2-butyne, the corresponding metallacycle was only obtained in modest yields (45 % for the reaction with 2-butyne) and a second major product, vinylimido complex 9, was formed in almost equal amounts (42 %) by 1,2-amination (formal insertion of the alkyne). The formation of compounds 7 a and 9 followed in part the same sequence of reaction steps and a key intermediate, an azirinido complex, represented a “bifurcation point” in the reaction network. Reaction of 1.2 equivalents of several diarylhydrazines and various substituted alkynes (1 equiv) at ambient temperature (or at 80 °C) in the presence of 10 mol % [Zr{(NXyl)2Npy}(NMe2)2] (2 a) gave the corresponding indole derivatives. On the other hand, the replacement of 1,1-diarylhydrazines by 1-methyl-1-phenyl hydrazine led to head-to-head cis-1,3-enynes in good yields.

Co-reporter:Dr. Susanne C. Martens;Dr. Ute Zschieschang;Dr. Hubert Wadepohl;Dr. Hagen Klauk;Dr. Lutz H. Gade
Chemistry - A European Journal 2012 Volume 18( Issue 12) pp:3498-3509
Publication Date(Web):
DOI:10.1002/chem.201103158

Abstract

A range of 2,9-perfluoroalkyl-substituted tetraazaperopyrene (TAPP) derivatives (15) was synthesised by reacting 4,9-diamino-3,10-perylenequinone diimine (DPDI) with the corresponding carboxylic acid chloride or anhydride in the presence of a base. The reaction of compounds 14 with dichloroisocyanuric acid (DIC) in concentrated sulphuric acid resulted in the fourfold substitution of the tetraazaperopyrene core, yielding the 2,9-bisperfluoroalkyl-4,7,11,14-tetrachloro-1,3,8,10-tetraazaperopyrenes 69, respectively. The optical and electrochemical data demonstrate the drastic influence of the core substitution on the properties. All compounds are highly luminescent (fluorescence quantum yields of up to Φ=0.8). The LUMO energies of the tetrachlorinated TAPP derivatives (determined by cyclic voltammetry and computed by DFT calulations) were found to be below −4 eV. In the course of this work the performance of TAPP derivatives in organic thin-film transistors (TFTs) was investigated, and their n-channel characteristics with field-effect mobilities of up to 0.14 cm2 V−1 s−1 and an on/off current ratio of >106 were confirmed. Long-term stabilities of 3–4 months under ambient conditions of the devices were established. Complementary inverters and ring oscillators with n-channel TFTs based on compound 8 and p-channel TFTs based on dinaphtho-[2,3-b:2′,3′-f]thieno[3,2-b]thiophene (DNTT) were fabricated on a glass substrate.

Co-reporter:Dr. Lara-Isabel Rodríguez;Dipl.-Chem. Torsten Roth;Dr. Julio LloretFillol;Dr. Hubert Wadepohl ;Dr. Lutz H. Gade
Chemistry - A European Journal 2012 Volume 18( Issue 12) pp:3721-3728
Publication Date(Web):
DOI:10.1002/chem.201103140

Abstract

A series of chiral mono-, di-, and trinuclear gold(I) complexes have been prepared and used as precatalysts in the asymmetric cyclohydroamination of N-protected γ-allenyl sulfonamides. The stereodirecting ligands were mono-, di-, and tridentate 2,5-diphenylphospholanes, which possessed C1, C2, and C3 symmetry, respectively, thereby rendering the catalytic sites in the di- and trinuclear complexes symmetry equivalent. The C3-symmetric trinuclear complex displayed the highest activity and enantioselectivity (up to 95 % ee), whilst its mono- and dinuclear counterparts exhibited considerably lower enantioselectivities and activities. A similar trend was observed in a series of mono-, di-, and trinuclear 2,5-dimethylphospholane gold(I) complexes. Aurophilic interactions were established from the solid-state structures of the trinuclear gold(I) complexes, thereby raising the question as to whether these secondary forces were responsible for the different catalytic behavior observed.

Co-reporter:Sonja Geib, Susanne C. Martens, Ute Zschieschang, Florian Lombeck, Hubert Wadepohl, Hagen Klauk, and Lutz H. Gade
The Journal of Organic Chemistry 2012 Volume 77(Issue 14) pp:6107-6116
Publication Date(Web):June 25, 2012
DOI:10.1021/jo300894p
A series of new tetraazapyrene (TAPy) derivatives has been synthesized by reducing 1,4,5,8-tetranitronaphthalene to its corresponding tin salt (I) and reacting it with perfluorinated alkyl or aryl anhydrides. The resulting 2,7-disubstituted TAPy molecules and the known parent compound 1,3,6,8-tetraazapyrene (II) have been further derivatized by core chlorination and bromination. The brominated compounds served as starting materials for Suzuki cross-coupling reactions with electron-poor arylboronic acids. Single-crystal X-ray analyses established polymorphism for some TAPy compounds. The ground-state geometries of all new TAPy derivatives were modeled with DFT methods [B3PW91/6-31 g(d,p) and B3PW91/6-311+g(d,p)], especially focusing on the energies of the lowest unoccupied molecular orbital (LUMO) and the electron affinities (EA) of the molecules. The results of the calculations were confirmed experimentally by cyclic voltammetry to evaluate the substitution effects at the 2 and 7 position and the core positions, respectively, and gave LUMO energy levels that range from −3.57 to −4.14 eV. Fabrication of organic field-effect transistors (OFETs) with several of these tetraazapyrenes established their potential as organic n-type semiconductors.
Co-reporter:Yan-Biao Kang
Journal of the American Chemical Society 2011 Volume 133(Issue 10) pp:3658-3667
Publication Date(Web):February 16, 2011
DOI:10.1021/ja110805b
Evidence for the protiocatalytic nature of the diacetoxylation of alkenes using PhI(OAc)2 as oxidant is presented. Systematic studies into the catalytic activity in the presence of proton-trapping and metal-complexing agents indicate that protons act as catalysts in the reaction. Using triflic acid as catalyst, the selectivity and reaction rate of the conversion is similar or superior to most efficient metal-based catalysts. Metal cations, such as Pd(II) and Cu(II), may interact with the oxidant in the initiation phase of the catalytic transformation; however, 1 equiv of strong acid is produced in the first cycle which then functions as the active catalyst. Based on a kinetic study as well as in situ mass spectrometry, a mechanistic cycle for the proton-catalyzed reaction, which is consistent with all experimental data presented in this work, is proposed.
Co-reporter:Matthias Kruck, Désirée C. Sauer, Markus Enders, Hubert Wadepohl and Lutz H. Gade  
Dalton Transactions 2011 vol. 40(Issue 40) pp:10406-10415
Publication Date(Web):20 Jun 2011
DOI:10.1039/C1DT10617A
Condensation of phthalodinitrile and 2-amino-5,6,7,8-tetrahydroquinoline gave the bis(2-pyridylimino)isoindole protioligand 1 (thqbpiH) in high yield. Deprotonation of thqbpiH (1) using LDA in THF at −78 °C yields the corresponding lithium complex [Li(THF)(thqbpi)] (2) in which the lithium atom enforces almost planar arrangement of the tridentate ligand, with an additional molecule of THF coordinated to Li. Reaction of cobalt(II) chloride or iron(II) chloride with one equivalent of the lithium complex 2 in THF led to formation of the metal complexes [CoCl(THF)(thqbpi)] (3a) and [FeCl(THF)(thqbpi)] (3b). The paramagnetic susceptibility of 3a,b in solution was measured by the Evans method (3a: μeff = 4.17 μB; 3b: μeff = 5.57 μB). Stirring a solution of 1 and cobalt(II) acetate tetrahydrate in methanol yielded the cobalt(II) complex 4 which was also accessible by treatment of 3a with one equivalent of silver or thallium acetate in DMSO. Whereas 3a,b were found to be mononuclear in the solid state, the acetate complex 4 was found to be dinuclear, the two metal centres being linked by an almost symmetrically bridging acetate. For all transition metal complexes paramagnetic 1H as well as 13C NMR spectra were recorded at variable temperatures. The complete assignment of the paramagnetic NMR spectra was achieved by computation of the spin densities within the complexes using DFT. The proton NMR spectra of 3a and 3b displayed dynamic behaviour. This was attributed to the exchange of coordinating solvent molecules by an associative mechanism which was analysed using lineshape analysis (ΔS≠= −154 ± 25 J mol−1 K−1 for 3a and ΔS≠ = −168 ± 15 J mol−1 K−1 for 3b).
Co-reporter:Thorsten Gehrmann;Dr. Julio LloretFillol;Solveig A. Scholl;Dr. Hubert Wadepohl ;Dr. Lutz H. Gade
Angewandte Chemie 2011 Volume 123( Issue 25) pp:5876-5881
Publication Date(Web):
DOI:10.1002/ange.201101070
Co-reporter:Susanne C. Martens, Till Riehm, Sonja Geib, Hubert Wadepohl, and Lutz H. Gade
The Journal of Organic Chemistry 2011 Volume 76(Issue 2) pp:609-617
Publication Date(Web):December 23, 2010
DOI:10.1021/jo102141w
A series of 2,9-bisaryl-1,3,8,10-tetraazaperopyrene (TAPP) derivatives has been synthesized by reacting 4,9-diamino-3,10-perylenequinone diimine with a large excess of the corresponding benzoyl chloride in refluxing nitrobenzene. Among all derivatives only ortho-substituted phenyl congeners were sufficiently soluble for studying solutions of defined concentration in organic solvents. The molecular structures of the crystallized compounds, determined by X-ray diffraction of four derivatives, are determined by the planar tetraazaperopyrene core and the interplanar angle of the phenyl rings, which depends on the size of the ortho substituent (40−70°). The intermolecular packing pattern of all compounds is characterized by parallel stacks of molecules with the substituted phenyl rings rotated out of the peropyrene plane to reduce the steric repulsion. Crystals of a TAPP derivative suitable for X-ray diffraction were grown from trifluoroacetic acid (TFA) for the first time, establishing a 2-fold protonated species. The ground-state geometries of the TAPP derivatives were calculated by DFT [B3PW91/6-31g(d,p)] and the lowest unoccupied molecular orbital (LUMO) energies of derivatives possessing electron-withdrawing groups were decreased, as were the computed electron affinities. The results of the modeling study were confirmed experimentally by cyclic voltammetry to evaluate the substituent effects on the highest occupied molecular orbital (HOMO) and the LUMO of the peropyrene core. The UV−vis absorption spectra of all compounds recorded in trifluoroacetic acid are almost superimposable and display a characteristic visible absorption band between 460 and 490 nm (log ε = 4.64−5.01) with a strong vibrational progression of 1173−1475 cm−1. Their fluorescence spectra are characterized by bands between 490 and 530 nm that are the mirror images of the absorption spectra (Stokes shifts of 10−50 nm). The luminescence quantum yields range from <0.01 to 0.30, thereby indicating a quenching effect for some substitution patterns.
Co-reporter:Dr. Qing-Hai Deng;Dr. Hubert Wadepohl ;Dr. Lutz H. Gade
Chemistry - A European Journal 2011 Volume 17( Issue 52) pp:14922-14928
Publication Date(Web):
DOI:10.1002/chem.201102375

Abstract

A new class of chiral tridentate N-donor pincer ligands, bis(oxazolinylmethylidene)isoindolines (boxmi), was synthesized in three steps starting from readily available phthalimides. Their reaction with ethyl (triphenylphosphoranylidene)acetate by means of a key-step Wittig reaction gave the ligand backbones, which were condensed with amino alcohols and then cyclized to obtain the corresponding ligands. These ligands were subsequently applied in the nickel(II)-catalyzed enantioselective fluorination of oxindoles and β-ketoesters to obtain the corresponding products with enantioselectivities of up to >99 % ee and high yields. Application of the chiral pincer ligands in the chromium-catalyzed enantioselective Nozaki–Hiyama–Kishi reaction of aldehydes gave the corresponding alcohols with an optimal enantioselectivity of 93 %.

Co-reporter:Dr. Julio LloretFillol;Achim Kruckenberg;Peter Scherl;Dr. Hubert Wadepohl ;Dr. Lutz H. Gade
Chemistry - A European Journal 2011 Volume 17( Issue 50) pp:14047-14062
Publication Date(Web):
DOI:10.1002/chem.201101864

Abstract

The modular one-pot synthesis of a large family of bi- and tridentate 2,5-dimethyl- and 2,5-diphenyl-substituted phospholanes employs air-stable chiral phospholanium chloride salts and primary amines or NH4Cl as starting materials. These were transformed into the C2-symmetric dimethyl- and diphenylphospholane ligands, which reacted with [Rh(cod)2]BF4 (cod=1,5-cyclooctadiene) to yield the rhodium complexes [Rh(L)(cod)]BF4 (L=bisphospholane ligands). The corresponding trisphospholane complexes, 11 and 12, were obtained in high yields (81 and 92 %, respectively), and fully characterised by NMR spectroscopy, mass spectrometry and elemental analysis. Whilst in the C3-symmetric complex 11, containing the tridentate 2,5-dimethylphospholane, the ligand is bound symmetrically, different coordination behaviour was found for the diphenyl-substituted complex 12, in which the coordination of only two of the three phospholane moieties to the metal centre was observed. A DFT study at the B3PW91 level established minimum energy structures consistent with experimental findings in solution and in the solid state. The non-coordinated phospholane unit present in 12 allowed further modification of the complex through the coordination of AuIX (X=Cl, C6F5 and tris(trifluoromethyl)phenyl (FMes)) fragments to the pendant phosphane. To investigate the potential of the new ligands, the enantioselective hydrogenation of a series of prochiral olefins as benchmark substrates, using isolated Rh complexes as catalysts, was studied. The substrates included methyl esters of three dehydro-α-acetamido acids and two itaconic acid derivatives. In general good to excellent enantioselectivities (of up to >99 % ee) were observed. Ligand backbone modification by coordination of bulky AuX substituents to the free phospholane unit in complex 12 led to an outstanding enhancement of the catalyst performance and there was a clear correlation between the properties of the complex periphery and the enantioselectivity.

Co-reporter:Thorsten Gehrmann;Dr. Julio LloretFillol;Solveig A. Scholl;Dr. Hubert Wadepohl ;Dr. Lutz H. Gade
Angewandte Chemie International Edition 2011 Volume 50( Issue 25) pp:5757-5761
Publication Date(Web):
DOI:10.1002/anie.201101070
Co-reporter:José A. Camerano, Christoph Sämann, Hubert Wadepohl, and Lutz H. Gade
Organometallics 2011 Volume 30(Issue 3) pp:379-382
Publication Date(Web):January 4, 2011
DOI:10.1021/om1010116
The reaction of the sodium salts of ligands 1a,b (1a = 1,3-bis(2-(5-(3,5-xylyl)pyridyl)imino)-5,6-dimethylisoindole, 1b = 1,3-bis(2-(4-tert-butylpyridyl)imino)-5,6-dimethylisoindole) with [Ir(μ-Cl)(COD)]2 (COD = cyclooctadiene) and [Ir(μ-Cl)(C2H4)2]2 afforded the corresponding isoindolato complexes [{BPI(1a,b)}IrI(COD)] (2a,b) and [{BPI(1a,b)}IrI(C2H4)2] (3a,b), respectively. The catalytic activity of the complexes 2a,b was tested in the epoxidation of a wide range of non-electron-rich olefins, using PPO (PPO = 3-phenyl-2-(phenylsulfonyl)-1,2-oxaziridine) as oxidizing agent, giving the corresponding epoxides in moderate to high yields.
Co-reporter:José A. Camerano, Anne-Sophie Rodrigues, Frank Rominger, Hubert Wadepohl, Lutz H. Gade
Journal of Organometallic Chemistry 2011 696(7) pp: 1425-1431
Publication Date(Web):
DOI:10.1016/j.jorganchem.2011.01.009
Co-reporter:Manfred Matena;Meike Stöhr Dr.;Till Riehm Dr.;Jonas Björk;Susanne Martens;MatthewS. Dyer Dr.;Mats Persson ;Jorge Lobo-Checa Dr.;Kathrin Müller Dr.;Mihaela Enache;Hubert Wadepohl ;Jörg Zegenhagen Dr.;ThomasA. Jung Dr.;LutzH. Gade Dr.
Chemistry - A European Journal 2010 Volume 16( Issue 7) pp:2079-2091
Publication Date(Web):
DOI:10.1002/chem.200902596

Abstract

The structural chemistry and reactivity of 1,3,8,10-tetraazaperopyrene (TAPP) on Cu(111) under ultra-high-vacuum (UHV) conditions has been studied by a combination of experimental techniques (scanning tunneling microscopy (STM) and X-ray photoelectron spectroscopy, XPS) and DFT calculations. Depending on the deposition conditions, TAPP forms three main assemblies, which result from initial submonolayer coverages based on different intermolecular interactions: a close-packed assembly similar to a projection of the bulk structure of TAPP, in which the molecules interact mainly through van der Waals (vDW) forces and weak hydrogen bonds; a porous copper surface coordination network; and covalently linked molecular chains. The Cu substrate is of crucial importance in determining the structures of the aggregates and available reaction channels on the surface, both in the formation of the porous network for which it provides the Cu atoms for surface metal coordination and in the covalent coupling of the TAPP molecules at elevated temperature. Apart from their role in the kinetics of surface transformations, the available metal adatoms may also profoundly influence the thermodynamics of transformations by coordination to the reaction product, as shown in this work for the case of the Cu-decorated covalent poly(TAPPCu) chains.

Co-reporter:Thorsten Gehrmann, Julio Lloret Fillol, Hubert Wadepohl and Lutz H. Gade
Organometallics 2010 Volume 29(Issue 1) pp:28-31
Publication Date(Web):December 14, 2009
DOI:10.1021/om900877c
Reaction of the zirconium complex [Zr(N2TBSNpy)(═NNPh2)(py)] with metal carbonyl complexes yielded the heteronuclear complexes [(N2TBSNpy)(py)Zr{μ-OCNNPh2-1κC:2κ2NO}Mm(CO)n] (M = Cr, Mo, W (m = 1, n = 5), Fe (m = 1, n = 4), Mn (m = 2, n = 9)). The bridging N-aminoisocyanato ligands in the dinuclear complexes were formed by [2 + 2] cycloaddition of the Zr═NNR2 unit and coordinated CO.
Co-reporter:JuttaK. Kassube ;LutzH. Gade
Advanced Synthesis & Catalysis 2009 Volume 351( Issue 5) pp:739-749
Publication Date(Web):
DOI:10.1002/adsc.200900016
Co-reporter:JuttaK. Kassube;Hubert Wadepohl ;LutzH. Gade
Advanced Synthesis & Catalysis 2009 Volume 351( Issue 4) pp:607-616
Publication Date(Web):
DOI:10.1002/adsc.200800741
Co-reporter:Felix Konrad ; Julio Lloret Fillol ; Hubert Wadepohl
Inorganic Chemistry 2009 Volume 48(Issue 17) pp:8523-8535
Publication Date(Web):August 12, 2009
DOI:10.1021/ic901189k
Bis(oxazolinylmethyl)pyrrole derivatives RLNH (4a−e), which were designed as protioligands for meridionally coordinating “pincer” ligands, were synthesized by cyclization of pyrrole-2,5-diethylacetate with a series of chiral amino alcohols. Deprotonation of RLNH (4a−d) with tBuLi and subsequent reaction with [RhCl(CO)2]2 gave the corresponding rhodium(I) complexes [Rh(RLN)(CO)] (R = iPr: 5a, Ph: 5b, Bn: 5c, Ind: 5d), which were also prepared by reaction of RLNH with one molar equivalent of [Rh(acac)(CO)2]. Upon heating the compounds at 100 °C in toluene over a period of 2−5 h, complete rearrangement via a 1,3-H shift between the pyrrole ring and the bridging methylene groups took place to yield the corresponding isomeric complexes [Rh(iso-RLN)(CO)] (6a−d). The transformation induced a planarization of the tridentate ligand system, resulting from the formation of a series of conjugated double bonds. Stirring the rhodium(I) complex 5a with an excess of CH3I in dichloromethane at ambient temperature14 gave the octahedrally coordinated product of an oxidative addition [Rh(iPrLN)(CH3)I(CO)] (7), while reaction of complex 5a with one molar equiv of CsBr3 as a mild brominating reagent in toluene at 80 °C led to complete conversion of the rhodium(I) species to the dibromorhodium(III) complex [Rh(iso-iPrLN)Br2(CO)] (8).
Co-reporter:Katharina Weitershaus, Benjamin D. Ward, Raphael Kubiak, Carsten Müller, Hubert Wadepohl, Sven Doye and Lutz H. Gade  
Dalton Transactions 2009 (Issue 23) pp:4586-4602
Publication Date(Web):15 Apr 2009
DOI:10.1039/B902038A
A series of new titanium half sandwich complexes, containing a 2-aminopyrrolinato ligand {NXylN}− as the ancillary ligand, have been prepared and are shown to be pre-catalysts for the hydroamination of alkynes. The coordination of {NXylN}− to titanium was achieved by reaction of [Cp*TiMe3] with the protioligand NXylNH giving [Cp*Ti(NXylN)(Me)2] (1). Upon reaction of complex 1 with an excess of tert-butylamine, the imido complex [Cp*Ti(NXylN)(NtBu)(NH2tBu)] (2) was formed. The latter provided the preparative entry to the synthesis of a range of N-aryl substituted imido complexes. Imido ligand exchange with 2,6-dimethylaniline, 2,4,6-trimethylaniline as well as 2,6-diisopropylaniline gave the corresponding arylimido complexes 3–5 in clean reactions. Reaction of the titanium imido complex [Cp*Ti(NXylN)(NtBu)(NH2tBu)] 2 with terminal arylacetylenes, such as phenylacetylene and tolylacetylene, led to C–H activation and the formation of alkynyl/amido complexes, whereas the arylimido complexes 3 and 5 cleanly underwent {2 + 2} cycloaddition, giving the azatitanacyclobutene derivatives. A single-crystal X-ray structure analysis of the azatitanacyclobutene [Cp*Ti(NXylN){κ2N(2,6-C6H3Me2)CTolCH}] (11) provided the first crystallographically characterized Markovnikov cycloaddition product of an imidotitanium complex with a terminal alkyne. The mechanistic aspects of the hydromanination of alkynes with the new Ti half sandwich complexes were studied and established a reversible {2 + 2} cycloaddition step and the cleavage of the metallacyclic intermediate as the rate determining step in the catalytic cycle. The titanium half sandwich imido complexes were found to be active catalysts for the inter- and intramolecular hydroamination of a broad range of alkynes and ω-aminoalkynes.
Co-reporter:Felix Konrad;Julio Lloret Fillol;Christoph Rettenmeier;Hubert Wadepohl
European Journal of Inorganic Chemistry 2009 Volume 2009( Issue 33) pp:4950-4961
Publication Date(Web):
DOI:10.1002/ejic.200900789

Abstract

The synthesis of the three ligands employed in this study is based on the condensation of two molar equivalents of (S)-valinol with the diester precursors pyrrole-2,5-bis(ethyl)acetate (2a), furan-2,5-bis(ethyl)acetate (2b) and thiophene-2,5-bis(ethyl)acetate (2c). This gave the corresponding bis(oxazolinylmethyl)pyrrole (iPrLNH, 4a), bis(oxazolinylmethyl)furan (iPrLO, 4b) and bis(oxazolinylmethyl)thiophene (iPrLS, 4c) ligands. Stirring 4a in MeOD at ambient temperature in the presence of a catalytic amount of acetic acid (1 mol-%) led to complete hydrogen/deuterium exchange in the two bridging methylene groups of the ligand. This behaviour is explained by an acetate-mediated reversible proton transfer between the oxazoline N atom and the methylene bridge, a conjecture which was supported by a DFT study of the process. Deprotonation of iPrLNH (4a) with tBuLi at –78 °C and subsequent stirring with NiCl2 yielded the square planar nickel(II) complex [Ni(iPrLN)Cl] (5). However, on stirring iPrLNH (4a) with nickel acetate in methanol, a deep red nickel acetato complex [Ni(iso-iPrLN)(OAc)] (6) bearing the isomerized tridentate pincer ligand was obtained. Reaction of acetato complex 6 with Me3SiCl in dichloromethane cleanly gave the corresponding chlorido complex [Ni(iso-iPrLN)Cl] (7), which is the isomer of compound 5. The intraligand rearrangement was explained by acetate-mediated proton transfer between the methylene bridges and the 3/4-positions of the pyrrole ring, the computed thermodynamic driving force being ΔG = –9.8 kcal mol–1. Neither thiophene derivative 4c nor furan-derived ligand 4b gave robust, isolable complexes with nickel(II). However, reaction of iPrLO (4b) with [CrCl3(thf)3] in thf yielded the yellow-green complex [CrCl3(iPrLO)] (8), whereas no complexation occurred with the analogous thiophene-derived bisoxazoline 4c. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2009)

Co-reporter:Benjamin D. Ward;Lenka Luke&x161;ová;Hubert Wadepohl;Stéphane Bellemin-Laponnaz
European Journal of Inorganic Chemistry 2009 Volume 2009( Issue 7) pp:866-871
Publication Date(Web):
DOI:10.1002/ejic.200801106

Abstract

The C3-symmetric trisoxazoline-supported scandium complex [Sc(iPr-trisox)(CH2SiMe3)3] (2) is highly active in the stereospecific polymerization of propene, 1-butene, 1-pentene, 1-hexene and 1-heptene, when activated with two equivalents of [Ph3C][B(C6F5)4]. The polymers thus produced were found to possess narrow molecular weight distributions and high levels of tacticity control (up to 99 % mmmm). Some insight into the nature of the active species was obtained by 1H, 13C and 29Si NMR experiments. In particular, the formation of two equivalents of Ph3CCH2SiMe3 at ambient temperature was observed alongside a C3-symmetric scandium complex tentatively assigned as the dication [Sc(iPr-trisox)(CH2SiMe3)]2+.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2009)

Co-reporter:Nathanaëlle Schneider;Matthias Kruck;Stéphane Bellemin-Laponnaz;Hubert Wadepohl
European Journal of Inorganic Chemistry 2009 Volume 2009( Issue 4) pp:493-500
Publication Date(Web):
DOI:10.1002/ejic.200801107

Abstract

A series of bidentate oxazoline-NHC ligands has been synthesized in which the two heterocycles are connected by a CH2 linker. The corresponding rhodium(I) complexes were prepared by direct deprotonation of the imidazolium halide salts followed by the addition of a solution of [Rh(nbd)Cl]2 at low temperature. The cationic square planar rhodium complexes were generated by halide abstraction via addition of an excess of KPF6 in a CH2Cl2/water solvent system. Alternatively, the deprotonation of the imidazolium hexafluorophosphates and reaction with [Rh(nbd)Cl]2 directly gave the complex cations. These as well as oxazoline-NHC systems, in which the two heterocycles are directly connected or through a CMe2 bridge, were investigated in the rhodium-catalyzed hydrosilylation of acetophenone. The comparison of the three ligand families showed that the catalysts obtained by direct coupling of oxazolines and N-heterocyclic carbenes, generating highly rigid chelate ligands, remain the most efficient systems giving the secondary alcohols in high enantioselectivity.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2009)

Co-reporter:Thorsten Gehrmann;Julio Lloret Fillol Dr.;Hubert Wadepohl Dr. ;LutzH. Gade Dr.
Angewandte Chemie International Edition 2009 Volume 48( Issue 12) pp:2152-2156
Publication Date(Web):
DOI:10.1002/anie.200805631
Co-reporter:Nathanaëlle Schneider Dr.;Markus Finger Dr.;Christian Haferkemper;Stéphane Bellemin-Laponnaz Dr.;Peter Hofmann Dr.;LutzH. Gade Dr.
Angewandte Chemie International Edition 2009 Volume 48( Issue 9) pp:1609-1613
Publication Date(Web):
DOI:10.1002/anie.200804993
Co-reporter:Thorsten Gehrmann;Julio Lloret Fillol Dr.;Hubert Wadepohl Dr. ;LutzH. Gade Dr.
Angewandte Chemie 2009 Volume 121( Issue 12) pp:2186-2190
Publication Date(Web):
DOI:10.1002/ange.200805631
Co-reporter:Katharina Weitershaus, Julio Lloret Fillol, Hubert Wadepohl and Lutz H. Gade
Organometallics 2009 Volume 28(Issue 16) pp:4747-4757
Publication Date(Web):July 23, 2009
DOI:10.1021/om900428d
Reaction of [Cp*Ti(NXylN){N-NPh2}(NH2tBu)] (1a) with 1 molar equiv of phenylallene resulted in the formation of a mixture of the two metallacyclic compounds [Cp*Ti(NXylN){κ2N(NPh2)C(CHPh)CH2}] (2a and 2b) in a molar ratio of 3:1. Comparison of the 13C and 15N NMR resonances with the computed chemical shifts, along with an X-ray diffraction study of 2a, allowed the assignment of the diastereomers as the cycloadducts derived from a (3,2)-cycloaddition of the allene to the Ti═N bond. Upon reaction of 1a with phenylisothiocyanate, two complexes were formed in a 3:1 ratio. The major isomer, [Cp*Ti(NXylN){κ2-N(NPh2)C(NPh)S}] (4a), was isolated and identified by X-ray diffraction to result from a [2+2] cycloaddition of the C═S bond in PhNCS to the Ti═N bond of the hydrazinediido unit, giving a four-membered metallacyclic ring with an S,N-coordination to the titanium center and an exocyclic C═N double bond. In the minor compound (4b) the two nitrogen atoms of the thiourea unit formed in the cycloaddition are coordinated to the titanium center. Isomerization between the cycloadducts occurs only in the direction 4b → 4a, and crossover experiments indicate a dissociative mechanism (via a retro-cycloaddition) for the rearrangement. The reaction of 1a with PhNCO gave analogous N,O- and N,N-coordinated cycloadducts. Finally, the N,N-dimethyl hydrazinediido complexes reacted unspecifically with most organic azides at ambient temperature. However, complex [Cp*Ti(NXylN){N-NMe2}(dmap)] (1c) underwent a clean and fast reaction with trimethylsilylazide to give the N-silylated η2-hydrazido(1−) titanium azide [Cp*Ti(NXylN)(η2-NMe2-NSiMe3)(N3)] (8), in which the trimethylsilyl group of the azide is transferred to the hydrazido ligand, while the azide is terminally coordinated to the titanium center.
Co-reporter:Katharina Weitershaus, Hubert Wadepohl and Lutz H. Gade
Organometallics 2009 Volume 28(Issue 12) pp:3381-3389
Publication Date(Web):May 12, 2009
DOI:10.1021/om900155t
Reaction of [Cp*Ti(NXylN)(NtBu)(NH2tBu)] with 1 molar equiv of diphenylhydrazine yielded the hydrazinediido complex [Cp*Ti(NXylN)(NNPh2)(NH2tBu)] (1a), whereas the orange pyridine adduct [Cp*Ti(NXylN)(NNPh2)(py)] (1b) was obtained by reacting the imide [Cp*Ti(NXylN)(NtBu)(NH2tBu)] with diphenylhydrazine in the presence of pyridine. The tert-butylamine coordinated to the metal center in 1a could be removed by heating the solid at 60 °C and 10−6 mbar for 72 h, yielding [Cp*Ti(NXylN)(NNPh2)] (1c). In the presence of pyridine or 4-dimethylaminopyridine (dmap) as neutral co-ligands the hydrazinediido complexes [Cp*Ti(NXylN)(NNMePh)(py)] (2a) and [Cp*Ti(NXylN)(NNMePh)(dmap)] (2b) as well as [Cp*Ti(NXylN)(NNMe2)(dmap)] (3) were prepared. Upon replacement of dmap by the weaker donor ligand pyridine in the synthesis of the pyridine adduct analogous to 3, a mixture of [Cp*Ti(NXylN)(NNMe2)(py)] (4a) and the dinuclear complex [Cp*2Ti2(NXylN)2(μ-η1,η1-NNMe2)(μ-η1,η2-NNMe2)] (4b) was obtained. Reaction of the dimethylhydrazinediido complex 3 with phenylacetylene gave the Markovnikov cycloadduct, which had sufficient lifetime to allow its 1H, 13C, and 15N NMR spectroscopic characterization in solution. All three hydrazinediido compounds 1a, 2a, and 3 were found to display remarkable activities in catalytic hydrohydrazinations at ambient temperatures. For catalyst loadings of 5 mol % complete conversions of the terminal alkynes and diynes with Markovnikov regioselectivities of over 99% selectivity were observed within 1 h.
Co-reporter:Nathanaëlle Schneider Dr.;Markus Finger Dr.;Christian Haferkemper;Stéphane Bellemin-Laponnaz Dr.;Peter Hofmann Dr.;LutzH. Gade Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 43) pp:11515-11529
Publication Date(Web):
DOI:10.1002/chem.200901594

Abstract

A detailed density functional theory (DFT) computational study (using the BP86/SV(P) and B3LYP/TZVP//BP86/SV(P) level of theory) of the rhodium-catalyzed hydrosilylation of ketones has shown three mechanistic pathways to be viable. They all involve the generation of a cationic complex [LnRhI]+ stabilized by the coordination of two ketone molecules and the subsequent oxidative addition of the silane, which results in the Rh–silyl intermediates [LnRhIII(H)SiHMe2]+. However, they differ in the following reaction steps: in two of them, insertion of the ketone into the RhSi bond occurs, as previously proposed by Ojima et al., or into the SiH bond, as proposed by Chan et al. for dihydrosilanes. The latter in particular is characterized by a very high activation barrier associated with the insertion of the ketone into the SiH bond, thereby making a new, third mechanistic pathway that involves the formation of a silylene intermediate more likely. This “silylene mechanism” was found to have the lowest activation barrier for the rate-determining step, the migration of a rhodium-bonded hydride to the ketone that is coordinated to the silylene ligand. This explains the previously reported rate enhancement for R2SiH2 compared to R3SiH as well as the inverse kinetic isotope effect (KIE) observed experimentally for the overall catalytic cycle because deuterium prefers to be located in the stronger bond, that is, CD versus MD.

Co-reporter:Manuela Gaab;Stéphane Bellemin-Laponnaz Dr.;LutzH. Gade
Chemistry - A European Journal 2009 Volume 15( Issue 22) pp:5450-5462
Publication Date(Web):
DOI:10.1002/chem.200900504
Co-reporter:Till Riehm, Gabriele De Paoli, Hubert Wadepohl, Luisa De Cola and Lutz H. Gade  
Chemical Communications 2008 (Issue 42) pp:5348-5350
Publication Date(Web):19 Sep 2008
DOI:10.1039/B811864G
A new class of N–B–N functionalized perylenes derived from N,N′,N″,N‴-diborylene-3,4,9,10-tetraaminoperylene (DIBOTAP) is readily accessible via several synthetic routes; they display intense green fluorescence and undergo two electrochemically reversible one-electron oxidation steps at ∼−0.2 V and 0.3 V (vs.SCE).
Co-reporter:JuttaK. Kassube;Hubert Wadepohl ;LutzH. Gade
Advanced Synthesis & Catalysis 2008 Volume 350( Issue 7-8) pp:1155-1162
Publication Date(Web):
DOI:10.1002/adsc.200700586

Abstract

Two series of peripherally functionalised dendrimers were synthesised based on poly(ethyleneimine) and poly(amidoamine) (PAMAM) dendrimers by reaction of (R)-6-N-[(γ-carboxyl)butanoyl]-aminomethyl-2,2′-bis(diphenylphosphino)-1,1′-binaphthyl (“Glutaroyl-Aminap”) with first to fifth generation PPI and zeroth to fourth generation PAMAM dendrimers using ethyl-N,N-dimethylcarbodiimide (EDC)/1-hydroxybenzotriazole as coupling reagents. All dendritic ligands were characterised by NMR spectroscopy, elemental analysis and, for generations G0–G3, MALDI-TOF mass spectrometry. The relationship between the size/generation of the dendritic ligands and their catalytic properties was established in the asymmetric hydrosilylation of acetophenone. It could be shown that (R)-6-aminomethyl-2,2′-bis(diphenylphosphino)-1,1′-binaphthyl (“H-Aminap”) can be attached to the dendrimers without any significant loss of catalytic selectivity compared to the monomeric ligand Benzoyl-Aminap. The selectivities observed in these reactions were higher than those obtained using unfunctionalised BINAP as ligand.

Co-reporter:Till Riehm, Hubert Wadepohl and Lutz H. Gade
Inorganic Chemistry 2008 Volume 47(Issue 24) pp:11467-11469
Publication Date(Web):November 14, 2008
DOI:10.1021/ic8017453
Metalation of 3,4,9,10-tetraaminoperylene with boron and aluminum complex fragments has given rise to a new class of highly fluorescent N−B−N- and N−Al−N-substituted dyes with fluorescence quantum yields of up to 82%.
Co-reporter:Michaela Kilian, Hubert Wadepohl and Lutz H. Gade  
Dalton Transactions 2008 (Issue 5) pp:582-584
Publication Date(Web):14 Nov 2007
DOI:10.1039/B716064J
Reaction of the triamido stannate MeSi{SiMe2N[(R)-CHMePh]}3SnLi (1) with 0.5 molar equivalent of [RhCl(olefin)2]2 (olefin = COE, C2H4) or [RhCl(PiPr3)2]2 yielded the Rh–Sn complexes [MeSi{SiMe2N[(R)-CHMePh]}2{SiMe2N[(R)-CHMe(η6-C6H5)}SnRh(L)] (L = COE: 2a, C2H4: 2b, PiPr33); their intramolecular η6-coordination, along with the tin–rhodium bond, represents the first “ansa” π-arene/stannate system.
Co-reporter:Heike Herrmann, Hubert Wadepohl and Lutz H. Gade  
Dalton Transactions 2008 (Issue 16) pp:2111-2119
Publication Date(Web):13 Mar 2008
DOI:10.1039/B800682B
The structures and properties of a series of new zirconium hydrazido(1−) complexes and the possibility of converting them to the respective hydrazido(2−) species are reported. Reaction of complex [Zr(N2TBSNpy)Cl2] (1) with the monolithiated hydrazide LiNHNMe2 gave the hydrazido(1−) complex [Zr(N2TBSNpy)(NHNMe2)Cl] (2) which exists as two isomeric forms (2a and 2b) in solution. All attempts to convert a mixture of 2a and 2b to the respective hydrazido(2−) compound by reaction with the bulky base lithium hexamethyldisilazide or via the alkyl/hydrazido(1−) complex [Zr(N2TBSNpy)(CH2SiMe3)(NHNMe2)] (3) and subsequent thermal alkane elimination failed. Reaction of 1 with LiHNNPhMe gave a mixture of stereoisomers of [Zr(N2TBSNpy)(NHNMePh)Cl] (4a and 4b), in which the hydrazido unit is end-on bound in solution and η2-bonded in the solid state. Reaction of this mixture with lithium hexamethyldisilazide in the presence of pyridine selectively yielded the hydrazido(2−) complex [Zr(N2TBSNpy)(NNPhMe)(py)] (5) which aggregated upon attempts to isolate it. Reaction of the insoluble precipitate with 4-dimethylaminopyridine (dmap) selectively gave the corresponding hydrazido(2−) complex [Zr(N2TBSNpy)(NNPhMe)(dmap)] (6), which could be obtained in a one-pot reaction directly from 1 and which was analytically and spectroscopically fully characterized. It appears that the isolation of stable hydrazido(2−) complexes of zirconium depends on the type of substituents at the Nβ atom as well as the co-ligands coordinated to the metal centre.
Co-reporter:Heike Herrmann, Thorsten Gehrmann, Hubert Wadepohl and Lutz H. Gade  
Dalton Transactions 2008 (Issue 44) pp:6231-6241
Publication Date(Web):30 Sep 2008
DOI:10.1039/B807808D
Reaction of the diamidozirconium complex [Zr(N2TBSNpy)(NMe2)2] (1) (N2TBSNpy = CH3C(C5H4N)(CH2NSiMe2tBu)2) or the diamidohafnium complex [Hf(N2TBSNpy)(NMe2)2] (2) with one molar equiv. of 1-aminopyridinium triflate in the presence of one equiv. of pyridine gave the corresponding (1-pyridinio)imido complexes [Zr(N2TBSNpy)(N–NC5H5)(OTf)(py)] (3) and [Hf(N2TBSNpy)(N–NC5H5)(OTf)(py)] (4). These were converted to the acetylide complexes [Zr(N2TBSNpy)(N–NC5H5)(CCPh)(py)] (5) and [Hf(N2TBSNpy)(N–NC5H5)(CCPh)(py)] (6) by reaction with lithium phenylacetylide and substitution of the triflato ligand. Upon reaction of 3 and 4 with one molar equivalent of R-NC (R = tBu, Cy, 2,6-xyl), N–N bond cleavage in the (1-pyridinio)imido unit took place and the respective carbodiimido complexes [M(N2TBSNpy](NCNR)(OTf)(py)] (7–12) were formed instantaneously. A similar type of reaction with CO gave the isocyanato complex [Zr(N2TBSNpy](NCO)(OTf)(py)] (13). Finally, the abstraction of the pyridine ligand in compounds 3 and 4 with B(C6F5)3 led to the formation of the triflato-bridged dinuclear complexes [Zr(N2TBSNpy)(N–NC5H5)(OTf)]2 (14) and [Hf(N2TBSNpy)(N–NC5H5)(OTf)]2 (15).
Co-reporter:Michaela Kilian;Hubert Wadepohl
European Journal of Inorganic Chemistry 2008 Volume 2008( Issue 11) pp:1892-1900
Publication Date(Web):
DOI:10.1002/ejic.200701240

Abstract

Reaction of the triamidostannates(II) MeSi{SiMe2N(p-tol)}3SnLi(OEt2) (1a) and MeSi{SiMe2N(3,5-xyl)}3SnLi(OEt2) (1b) with 1/2 molar equiv. of [RhCl(diolefin)]2 (diolefin = COD, NBD) and phosphanes, phosphites or isonitriles (“L”) gave the square-planar complexes [MeSi{SiMe2NAryl}3SnRh(L)(diolefin)] (Aryl = 3,5-xyl, p-tol) in which thestannates are directly bonded to rhodium through Rh–Sn bonds. In contrast, the analogous transformation of 1b with 1/2 equiv. of [RhCl(C2H4)2]2 and PiPr3 in toluene did not lead to a square-planar bisethylene complex [MeSi{SiMe2N(3,5-xyl)}3SnRh(C2H4)2(PiPr3)] but the brown 18e π-arene complex [MeSi{SiMe2N(3,5-xyl)}3SnRh(PiPr3)(η6-toluene)] (6). Equilibria of the square planar diolefin complexes with π-arene systems such as 6 were observed upon their dissolution in toluene.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2008)

Co-reporter:BjörnK. Langlotz;Hubert Wadepohl Dr. ;LutzH. Gade Dr.
Angewandte Chemie 2008 Volume 120( Issue 25) pp:4748-4752
Publication Date(Web):
DOI:10.1002/ange.200801150
Co-reporter:Manfred Matena;Till Riehm;Meike Stöhr Dr.;ThomasA. Jung Dr.;LutzH. Gade Dr.
Angewandte Chemie International Edition 2008 Volume 47( Issue 13) pp:2414-2417
Publication Date(Web):
DOI:10.1002/anie.200704072
Co-reporter:BjörnK. Langlotz;Hubert Wadepohl Dr. ;LutzH. Gade Dr.
Angewandte Chemie International Edition 2008 Volume 47( Issue 25) pp:4670-4674
Publication Date(Web):
DOI:10.1002/anie.200801150
Co-reporter:Béatrice Jacques Dr.;Clémence Dro Dr.;Stéphane Bellemin-Laponnaz Dr.;Hubert Wadepohl Dr.;LutzH. Gade Dr.
Angewandte Chemie International Edition 2008 Volume 47( Issue 24) pp:4546-4550
Publication Date(Web):
DOI:10.1002/anie.200800870
Co-reporter:BjörnK. Langlotz;Julio LloretFillol Dr.;JürgenH. Gross Dr.;Hubert Wadepohl Dr.;LutzH. Gade Dr.
Chemistry - A European Journal 2008 Volume 14( Issue 33) pp:10267-10279
Publication Date(Web):
DOI:10.1002/chem.200801373

Abstract

A new type of mediator for cobalt(II)-mediated radical polymerization is reported which is based on 1,3-bis(2-pyridylimino)isoindolate (bpi) as ancillary ligand. The modular synthesis of the bis(pyridylimino)isoindoles (bpiH) employed in this work is based on the condensation of 2-aminopyridines with phthalodinitriles. Reaction of the bpiH protio-ligands with a twofold excess of cobalt(II) acetate or cobalt(II) acetylacetonate in methanol gave [Co(bpi)(OAc)], which crystallize as coordination polymers, and a series of [Co(acac)(bpi)(MeOH)], which are mononuclear octahedral complexes. Upon heating the [Co(acac)(bpi)(MeOH)] compounds to 100 °C under high vacuum, the coordinated methanol was removed to give the five-coordinate complexes [Co(acac)(bpi)]. The polymerization of methyl acrylate at 60 °C was investigated by using one molar equivalent of the relatively short-lived radical source 2,2′-azobis(4-methoxy-2,4-dimethylvaleronitrile) (V-70) as initiator (monomer/catalyst/V-70: 600:1:1). The low solubility of the acetato complexes inhibits their significant activity as mediators in this reaction, whereas the acetylacetonate complexes control the radical polymerization of methyl acrylate more effectively. The radical polymerizations of the hexacoordinate complexes did not show a linear increase in number-average molecular weight (Mn) with conversion; however, the polydispersities were relatively low (PDI=1.12–1.40). By using the pentacoordinate complexes [Co(acac)(bpi)] as mediators, a linear increase in Mn values with conversion, which were very close to the theoretical values for living systems, and very low polydispersities (PDI<1.13) were obtained. This was also achieved in the block copolymerization of methyl acrylate and n-butyl acrylate. The intermediates with the growing acrylate polymer radical (.PA) were identified by liquid injection field desorption/ionization mass spectrometry as following the general formula [Co(acac)(4-methoxy-bpi)-(MA)n-R] (MA: methyl acrylate; R: C(CH3)(CH2C(CH3)2OCH3)CN), a notion also confirmed by NMR end-group analysis.

Co-reporter:Heike Herrmann Dr.;JulioLloret Fillol Dr.;Thorsten Gehrmann;Markus Enders Dr.;Hubert Wadepohl Dr. ;LutzH. Gade Dr.
Chemistry - A European Journal 2008 Volume 14( Issue 27) pp:8131-8146
Publication Date(Web):
DOI:10.1002/chem.200800876

Abstract

Reaction of the dichloro complexes [M(N2TBSNpy)Cl2] (M=Zr: 1, Hf: 2; TBS: tBuMe2Si; py: pyridine) with one molar equivalent of LiNHNPh2 gave mixtures of the two diastereomeric chlorohydrazido(1−) complexes [M(N2TBSNpy)(NHNPh2)Cl] (M=Zr: 3 a,b, Hf: 4 a,b) in which the diphenylhydrazido(1−) ligand adopts a bent κ1 coordination. This mixture of isomers could be cleanly converted into the deep green diphenylhydrazido(2−) complexes [Zr(N2TBSNpy)(NNPh2)(py)] (5) and [Hf(N2TBSNpy)(NNPh2)(py)] (6), respectively, by dehydrohalogenation with lithium hexamethyldisilazide (LiHMDS) in the presence of one molar equivalent of pyridine. Both complexes contain a linearly coordinated hydrazinediide for which a DFT-based frontier orbital analysis established bonding through one σ and two π orbitals. A high polarity of the MN bond was found, in accordance with the description of hydrazinediide(2−) acting as a six-electron donor ligand. The pyridine ligand in [M(N2TBSNpy)(NNPh2)(py)] (M=Zr: 5, Hf: 6) is substitutionally labile as established by line-shape analysis of the dynamic spectra (ΔG=19 kcal mol−1). A change in denticity of the hydrazido unit from κ1 to κ2 was studied by DFT methods. Both forms are calculated to be very close in energy and are only separated by shallow activation barriers, which supports the notion of a rapid κ1 to κ2 interconversion. This process is believed to happen early on in the NN scission in the presence of coupling reagents. Frontier orbital and natural population analyses suggest that a primarily charge-controlled nucleophilic attack at Nα is unlikely whereas interaction with an electrophile could play an important role. This hypothesis was tested by the reaction of 5 and 6 with one molar equivalent of B(C6F5)3 to give [Zr(N2TBSNpy)(NNPh2){B(C6F5)3}] (7) and [Hf(N2TBSNpy)(NNPh2){B(C6F5)3}] (8). In these products, B(C6F5)3 becomes attached to the Nα atom of the side-on bound hydrazinediide and there is an additional interaction of an ortho-F atom of a C6F5 ring with the metal centre.

Co-reporter:LutzH. Gade Dr.;Stéphane Bellemin-Laponnaz Dr.
Chemistry - A European Journal 2008 Volume 14( Issue 14) pp:4142-4152
Publication Date(Web):
DOI:10.1002/chem.200701990

Abstract

Rotational molecular symmetry, modularity and other aspects of ligand design have played a role in the development of a new class of stereodirecting ligands. The use of highly symmetrical, stereodirecting ligands may reduce the number of transition states and diastereomeric reaction intermediates and, in favourable cases, this degeneration of alternative reaction pathways may lead to high stereoselectivity in catalytic reactions and greatly simplifies the analysis of such transformations. In this concept article, we describe the way in which these considerations have played a role in the development of a new class of stereodirecting ligands. Tris(oxazolinyl)ethanes (“trisox”) have proved to be versatile ligand systems for the development of enantioselective catalysts of the d- and f-block metals employed in a wide range of catalytic conversions. These include Lewis acid catalysed transesterifications, CC and CN coupling reactions, the catalytic polymerisation of α-olefins as well as Pd-catalysed allylic alkylations. An overview of the current state of this field is given and the potential for further development will be highlighted.

Co-reporter:Nadia Vujkovic, Julio Lloret Fillol, Benjamin D. Ward, Hubert Wadepohl, Philip Mountford and Lutz H. Gade
Organometallics 2008 Volume 27(Issue 11) pp:2518-2528
Publication Date(Web):April 30, 2008
DOI:10.1021/om8001203
Reaction of the titanium imido complexes [Ti(NR)(N2XylNpy)(L)] (N2XylNpy = MeC(2-C5H4N){CH2N(3,5-C6H3Me2)}2) with aryl acetylenes afforded the {2+2} cycloaddition products [Ti(N2XylNpy){κ2-N(tBu)CH═CAr}] (Ar = Ph: 1a, Tol: 1b (Tol = 4-C6H4Me)). Complex 1a was found to insert sulfur and selenium atoms into the Ti−C bond of the strained azatitanacyclobutene unit to give the five-membered metallacyclic systems [Ti(N2XylNpy){κ2-N(tBu)CH═C(Ph)E}] (E = S: 2, Se: 3). Likewise, isonitriles were found to insert into the Ti−C bonds to give [Ti(N2XylNpy){κ2-N(tBu)CH═C(Ar)C═NR}] (Ar = Ph, R = xylyl: 4, Cy: 5, tBu: 6a; Ar = C6H5Me, R = tBu: 6b), one example of which (4) has been characterized by X-ray crystallography. The latter species is thought to be a key intermediate in the titanium-mediated three-component coupling of primary amines with acetylenes and isonitriles (“iminoamination”). Reaction of [Ti(N2XylNpy){κ2-N(tBu)CH═C(Ph)C═NR}] with tBuNH2 gave tBuN═CHC(Ph)CHNHR, which have been previously identified as the major reaction products in the catalytic aminoamination of alkynes.
Co-reporter:Michaela Kilian, Hubert Wadepohl and Lutz H. Gade
Organometallics 2008 Volume 27(Issue 4) pp:524-533
Publication Date(Web):January 19, 2008
DOI:10.1021/om700960p
Reaction of the triamidostannates(II) MeSi[SiMe2N(3,5-xyl)]3SnLi(OEt2) (2a) and MeSi[SiMe2N(p-tol)]3SnLi(OEt2) (2b) with 0.5 molar equiv of [RhCl(COD)]2 gave the zwitterionic complexes [MeSi[SiMe2NAryl]2Sn[SiMe2N(η6-Aryl)]Rh(diolefin)] (Aryl = 3,5-xyl: 3a, p-tol: 3b). In these one of the aryl groups acts as a η6-ligand, thus resulting in the 18-electron rhodium species. Addition of 1 equiv of PPh3 to a solution of 3a or 3b yielded the square-planar complexes [MeSi[SiMe2NAryl]3SnRh(PPh3)(COD)] (Aryl = 3,5-xyl: 4a, p-tol: 4b), in which the stannates are directly bonded to rhodium through Rh−Sn bonds. Treatment of the complexes 4a,b and 5a,b with hydrogen gas in the presence of benzene leads to the hydrogenation of the diolefin and its replacement by benzene as a formal six-electron donor ligand. These 18-electron complexes [MeSi[SiMe2N(p-tol)]3SnRh(PPh3)(η6-arene)] 6a,b and 7a,b are also accessible by reacting the stannates with 0.5 equiv of [RhCl(C2H4)2]2, PPh3, and the appropriate arene. Upon reacting the xylyl stannate 2a with [IrCl(COD)]2 and Ph3P, it was possible to isolate the square-planar Ir complex [MeSi[SiMe2N(3,5-xyl)]3SnIr(PPh3)(COD)] (8a). In contrast, for the tolyl stannate CH-activation occurred to give the IrIII compound [MeSi[SiMe2N(p-tol)]2[SiMe2N(2-C6H3-4-CH3)]SnIr(H)(PPh3)(COD)] (8b).
Co-reporter:Heike Herrmann, Julio Lloret Fillol, Hubert Wadepohl and Lutz H. Gade
Organometallics 2008 Volume 27(Issue 2) pp:172-174
Publication Date(Web):January 3, 2008
DOI:10.1021/om700955p
Reaction of a (1-pyridinio)imidozirconium complex with isocyanides or CO leads to instantaneous N−N bond cleavage and the concomitant formation of the carbodiimido complexes [Zr(N2TBSNpy](N═C═NR)(OTf)(py)] or the isocyanato complex [Zr(N2TBSNpy](NCO)(OTf)(py)], respectively.
Co-reporter:Lutz H. Gade, Stéphane Bellemin-Laponnaz
Coordination Chemistry Reviews 2007 Volume 251(5–6) pp:718-725
Publication Date(Web):March 2007
DOI:10.1016/j.ccr.2006.05.015
Whereas initially the use of chiral N-heterocyclic carbene (NHC) ligands in asymmetric catalysis met with very limited success, several novel structural concepts have emerged which have led to a rapid expansion of the field since the beginning of this decade. One concept in the design of new chiral NHC-containing molecular catalysts is the incorporation of oxazolines. These may play a secondary structural role in the design of monodentate chiral NHCs which are either employed as organocatalysts or as stereodirecting spectator ligands in transition metal complexes. Alternatively, oxazolines may be ligating units which are combined with NHCs to give polydentate stereodirecting ligands. Both approaches have been applied successfully and are reviewed in this article.
Co-reporter:Markus Wahl, Meike Stöhr, Hannes Spillmann, Thomas A. Jung and Lutz H. Gade  
Chemical Communications 2007 (Issue 13) pp:1349-1351
Publication Date(Web):27 Feb 2007
DOI:10.1039/B700909G
Fourfold symmetric zinc-octaethylporphyrin (OEP) has been incorporated in the holes of the hexagonal molecular network generated by thermal dehydrogenation of 4,9-diaminoperylene-quinone-3,10-diimine (DPDI) on a Cu(111) surface and displayed hindered rotation; the reorganization between the potential minima, a rotation–libration, which is characterized by an activation energy of ED = 0.17 ± 0.03 eV, has been monitored in the STM tunnelling currents as a bi-state “switching”.
Co-reporter:Lenka Lukešová, Benjamin D. Ward, Stéphane Bellemin-Laponnaz, Hubert Wadepohl and Lutz H. Gade  
Dalton Transactions 2007 (Issue 9) pp:920-922
Publication Date(Web):18 Jan 2007
DOI:10.1039/B700269F
The thulium complexes [Tm(iPr-trisox)(CH2SiMe2R)3] (R = Me 1a, Ph 1b) were synthesized from the thulium trialkyl precursors [Tm(CH2SiMe2R)3(thf)2]; reaction of 1a with two equivalents of [Ph3C][B(C6F5)4] gave a cationic complex 1c, which was found to polymerize 1-hexene, 1-heptene and 1-octene to give the corresponding polyolefins with moderate to good activities and with minimum isotacticity of 90%, 83% and 95%, respectively.
Co-reporter:Lutz H. Gade  Dr.;Guido Marconi Dr.;Clémence Dro;Benjamin D. Ward Dr.;Macarena Poyatos Dr.;Stéphane Bellemin-Laponnaz Dr.;Hubert Wadepohl  Dr.;Lorenzo Sorace Dr.;Giordano Poneti Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 11) pp:
Publication Date(Web):15 FEB 2007
DOI:10.1002/chem.200601651

A key feature of tris(oxazolinyl)ethane (“trisox”) ligands, which have shown broad scope in asymmetric catalysis, is the orientation and steric demand of their oxazoline substituents. This, along with the modularity of their synthesis determines their coordination chemistry. The possibility to combine oxazolines, in which the stereogenic centers adjacent to the N-donor atoms have different absolute configuration, whilst retaining their ability to coordinate as tripodal ligands, has been demonstrated by the synthesis of the enantiomerically pure C3-symmetric iPr-trisox(S,S,S) and C1-symmetric iPr-trisox(S,S,R) and their reaction with [Mo(CO)3(NCMe)3] yielding [Mo{iPr-trisox(S,S,S)}(CO)3] (1 a) and [Mo{iPr-trisox(S,S,R)}(CO)3] (1 b), respectively. The non-autocomplementarity of two homochiral trisox ligands at one metal center has been demonstrated by reaction of rac-C3iPr-trisox with one equivalent of [Co(ClO4)2]⋅6 H2O, giving the centrosymmetric heterochiral complex [Co(iPr-trisox)2](ClO4)2 (3), whereas an analogous reaction with the enantiopure ligand yielded a mixture of CoII complexes, which is characterized by the total absence of a [(trisox)2Co]+/2+ ion. The scope of the trisox ligand in terms of facial coordination to both early and late transition metals was demonstrated by the synthesis and structural characterization of the mononuclear complexes [ScCl3(iPr-trisox)] (4), [Fe(tBu-trisox)(NCMe)3](BF4)2 (5), and [Ru(η6-p-cymene)(iPr-trisox)](PF6)2 (6). The facial coordination of their three ligating atoms to a metal center may be impeded if the transition-metal center stereoelectronically strongly favors a non-deltahedral coordination sphere, which is generally the case for the heavier d8-transition-metal atoms/ions. Reaction of iPr-trisox with [Rh(cod)2]BF4 led to the formation of the 16-electron d8-configured complex [Rh(iPr-trisox)(cod)](BF4) (7), which is oxidized by CsBr3 to give the RhIII complex [RhBr3(iPr-trisox)] (8) possessing a C3-symmetric structure with a κ3-N-trisox ligand. The crystalline salts [M2(μ-Cl3)(iPr-trisox)2](PF6) (M=FeII: 9, CoII: 10, NiII: 11), were prepared by addition of one molar equivalent of iPr-trisox and an excess of KPF6 to solutions of the anhydrous (FeCl2) or hydrated metal halides (CoCl2⋅6 H2O, NiCl2⋅6 H2O). All dinuclear complexes display weak magnetic coupling. For the mononuclear species [CuCl2(iPr-trisox)] (12) the removal of a chloride anion and thus the generation of a dinuclear chloro-bridged structure failed due to Jahn–Teller destabilization of a potential octahedral coordination sphere.

Co-reporter:Carole Foltz;Markus Enders Priv.-Doz. Dr.;Stéphane Bellemin-Laponnaz Dr.;Hubert Wadepohl  Dr.;Lutz H. Gade  Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 21) pp:
Publication Date(Web):24 MAY 2007
DOI:10.1002/chem.200700307

Threefold symmetrical chiral podands may simplify the stereochemistry of key catalytic intermediates for cases in which they only act as bidentate ligands. This applies to systems in which chemical exchange between the different κ2-coordinated forms takes place and in which the non-coordinated sidearm may play a direct or indirect role at some earlier or later stage in the catalytic cycle. Palladium(II)-catalysed allylic substitutions provide appropriate test reactions along these lines. A series of neutral dichloropalladium(II) complexes, [PdCl2(iPr-trisox)] (1 a), [PdCl2(Ph-trisox)] (1 b), [PdCl2(Bn-trisox)] (1 c) and [PdCl2(Ind-trisox)] (1 d) (trisox=1,1,1-tris(oxazolinyl)ethane) were synthesised by reaction of the respective trisox derivative with [PdCl2(PhCN)2] and characterised inter alia by 15N NMR spectroscopy. Direct detection of the heteronuclei without isotope enrichment and with “normal” sample concentrations was achieved with the aid of a cryogenically cooled NMR probe on a 600 MHz NMR spectrometer. Whereas the 15N nuclei of the coordinated oxazoline rings resonate at δ=160–167 ppm and appear as two singlets due to their diastereotopicity, the signal assigned to the dangling oxazoline “arm” is observed at δ=238–240 ppm. Variable-temperature NMR studies along with a systematic series of magnetisation transfer experiments established exchange between ligating and non-ligating oxazoline rings. Reaction of [Pd(allyl)(cod)]BF4 (cod=cyclooctadiene) with Ph-trisox in CH2Cl2 gave the corresponding allyl complex 2, for which fast exchange between the three oxazoline heterocycles as well as between the exo and endo diastereomers was observed along with a very slow η313 process of the allyl fragment (magnetisation transfer). Palladium(0) complexes were prepared by reaction of trisox derivatives or sidearm-functionalised BOX (BOX=bis(oxazolinyl)dimethylmethane) ligands with [Pd(nbd)(alkene)] (nbd=norbornadiene, alkene=maleic anhydride or tetracyanoethylene). X-ray diffraction studies of the iPr-trisox and Ph-trisox complexes (3 a and 3 b) established Y-shaped trigonal planar coordination geometries with the trisox ligand coordinated in a bidentate fashion, whilst the π-coordinated maleic anhydride ligand adopts one of the two possible diastereotopic orientations. As the catalytic test reaction, the allylic alkylation of 1,3-diphenylprop-2-enyl acetate substrate with dimethyl malonate as nucleophile (in the presence of N,O-bis(trimethylsilyl)acetamide) was investigated for the trisox derivatives, their BOX analogues, and a series of less symmetric “sidearm” functionalised bisoxazolines. The trisoxazoline-based catalysts generally induce a better enantioselectivity compared to their bisoxazoline analogues and display significant reduction of the induction period as well as rate enhancement.

Co-reporter:Carole Foltz;Markus Enders Priv.-Doz. Dr.;Stéphane Bellemin-Laponnaz Dr.;Hubert Wadepohl  Dr.;Lutz H. Gade  Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 21) pp:
Publication Date(Web):5 JUL 2007
DOI:10.1002/chem.200790074

The triskelion, an ancient symbol possessing threefold rotational symmetry, is found in Celtic and Mediterranean cultures. It appears on the flags of the Isle of Man and Sicily, but also on many coats-of-arms throughout Europe. Dynamic exchange between three symmetry-equivalent binding sites is observed for palladium precatalysts that possess C3-chiral stereodirecting ligands. In their Full Paper on page 5994 ff., L. H. Gade, S. Bellemin-Laponnaz et al. describe such enantioselective catalysts bearing chiral tripods that display superior performance compared to their bidentate analogues.

Co-reporter:Till Riehm;Gabriele De Paoli;Asgeir E. Konradsson;Luisa De Cola  Dr.;Hubert Wadepohl  Dr.;Lutz H. Gade  Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 26) pp:
Publication Date(Web):19 JUN 2007
DOI:10.1002/chem.200700383

Tetraazaperopyrene and a range of derivatives have been synthesised and their photophysical and redox-chemical properties studied. The parent compound, 1,3,8,10-tetraazaperopyrene (1), was prepared by treating 4,9-diamino-3,10-perylenequinone diimine with triethyl orthoformate, whereas the 2,9-disubstituted derivatives of 1 were obtained after treatment with the corresponding carboxylic acid chloride or anhydride (2 mol equiv). The 1,3,8,10-tetraazaperopyrene core structure was established by X-ray diffraction of 2,9-bis(2-bromophenyl)-1,3,8,10-tetraazaperopyrene (6). The UV-visible absorption spectra of the compounds have a characteristic visible π*π absorption band at 440 nm (log εmax=4.80) with a strong vibrational progression (Δν≈1450 cm−1). Diprotonation of the nitrogen atoms induces a bathochromic shift of this band from 430–440 to 470–480 nm and all four nitrogen atoms are protonated when pure H2SO4 is used as the solvent. The first and second as well as the third and fourth protonations occur concomitantly, which implies that they have very similar pKa values and, consequently, similar proton affinities. A theoretical study of the proton affinities in the gas phase and in solution attributes this behaviour to the effects of polar solvents, which dampen the charge of a protonated site at the other end of the molecule and thus effectively decouple the two opposite pyrimidine units in the polycondensed aromatic compound. The photophysical data were modelled in a time-dependent DFT study of 1, 1H22+ and 1H44+ in both the gas phase and in a polar solvent. All the dyes show weak fluorescence in organic solvents, however, their protonated conjugate acids show dramatically increased fluorescence intensity. All of the dyes undergo two electrochemically reversible one-electron reductions with cyclovoltammetric half-wave potentials at Ered1≈−0.9 V and Ered2≈−1.3 V (vs. SCE), which are associated with characteristic spectral changes.

Co-reporter:Carole Foltz;Björn Stecker Dr.;Guido Marconi Dr.;Stéphane Bellemin-Laponnaz Dr.;Hubert Wadepohl  Dr.;Lutz H. Gade  Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 35) pp:
Publication Date(Web):23 OCT 2007
DOI:10.1002/chem.200701085

The underlying conceptual differences in exploiting two- and threefold rotational symmetry in the design of chiral ligands for asymmetric catalysis have been addressed in a comparative study of the catalytic performance of bisoxazoline (BOX) and tris(oxazolinyl)ethanes (trisox) containing copper(II) Lewis acid catalysts. The differences become apparent in constructing new catalysts by systematically “deforming” the stereodirecting ligand by inverting chiral centres or replacing chiral by achiral oxazolines. The catalytic α-amination of ethyl 2-methylacetoacetate with dibenzyl azodicaboxylate, which occurs with high enantioselectivity for both Ph2-BOX and Ph3-trisox copper catalysts, has been employed as the test reaction. In the trisox–copper complex [CuII(iPr3-trisox)(κ2-O,O′-MeCOCHCOOEt)]+[BF4] (1), which was characterised by X-ray diffraction, two of the oxazoline groups are coordinated to the central copper atom, whilst the third oxazoline unit is dangling with the N-donor pointing away from the metal centre. A similar arrangement is found for the stereochemically “mixed” C1-trisox complex [CuII{(Ph3-trisox(R,S,S)}(κ2-O,O′-MeCOCHCOOEt)(H2O)]+[ClO4] (2), in which the phenyl substituents adopt a first coordination sphere meso arrangement. The almost identical selectivity of the Ph3-trisox(R,R,R)- and Ph2-BOX(R,R)-derived catalysts is as expected from the proposed model of the active catalyst based on a partially decoordinated podand. The behaviour of the “desymmetrised” trisox–Cu catalysts may be rationalised in terms of a general steady-state kinetic model for the three possible active bisoxazoline–copper species, which are expected to be in rapid exchange with each other in solution. This applies to both the trisox derivatives with stereochemically inverted and achiral oxazoline rings. The study underscores previous observations of superior performance of the catalysts bearing C3-chiral stereodirecting ligands as compared to systems of lower symmetry.

Co-reporter:Heike Herrmann;Julio Lloret Fillol Dr.;Hubert Wadepohl  Dr.;Lutz H. Gade  Dr.
Angewandte Chemie International Edition 2007 Volume 46(Issue 44) pp:
Publication Date(Web):8 OCT 2007
DOI:10.1002/anie.200703938

Piece by piece: Hydrazides at zirconium centers undergo facile NN bond cleavage and react as equivalents for metallanitrenes, thus allowing the atom-by-atom assembly of bridging dinitridosulfate(IV) and dinitridoselenate(IV) ligands (see scheme; TBS=tBuMe2Si).

Co-reporter:Heike Herrmann;Julio Lloret Fillol Dr.;Hubert Wadepohl  Dr.;Lutz H. Gade  Dr.
Angewandte Chemie 2007 Volume 119(Issue 44) pp:
Publication Date(Web):8 OCT 2007
DOI:10.1002/ange.200703938

Stück für Stück: Die N-N-Bindungen von Hydraziden an Zirconiumzentren werden leicht gespalten. Diese Spezies reagieren daher wie Metallanitrene, sodass verbrückende Dinitridosulfat(IV)- und Dinitridoselenat(IV)-Liganden Atom für Atom aufgebaut werden können (siehe Schema; TBS=tBuMe2Si).

Co-reporter:Markus Seitz, Carmine Capacchione, Stéphane Bellemin-Laponnaz, Hubert Wadepohl, Benjamin D. Ward and Lutz H. Gade  
Dalton Transactions 2006 (Issue 1) pp:193-202
Publication Date(Web):14 Nov 2005
DOI:10.1039/B512570G
A series of sidearm functionalized bisoxazoline ligands has been synthesized by reaction of the monolithiated methyl{bis(oxazolinyl)}methane with the appropriate electrophiles, and tested in the copper catalyzed asymmetric allylic oxidation of cyclohexene (“Kharasch–Sosnovski” reaction). The observed enantioselectivities were higher (up to 85% ee) than for the unfunctionalized bisoxazoline (“BOX”) derivatives (ca. 60% ee). Regardless of the functional groups incorporated into the sidearm unit, the ee's obtained for the different derivatives were essentially indistinguishable. This implies that the sidearms do not interfere directly in this reaction and only play an indirect role by virtue of their steric demand. Three of the copper complexes have been characterized by X-ray diffraction, establishing a distorted octahedral coordination geometry around the copper atom in all three cases. In the elongated distorted CuN2O4 octahedra, the two nitrogen atoms of the oxazolines and one oxygen atom of each acetate ligand occupy the ‘equatorial’ positions whereas the sidearms do not interact with the metal centres.
Co-reporter:Massimiliano Forcato;Fredrik Lake;Myriam Mba Blazquez;Patrick Renner;Marco Crisma;Giulia Licini;Christina Moberg
European Journal of Inorganic Chemistry 2006 Volume 2006(Issue 5) pp:
Publication Date(Web):3 JAN 2006
DOI:10.1002/ejic.200500657

AlIII and TiIV complexes of C3-symmetric tetradentate trisamidoamine ligands with trigonal bipyramidal coordination geometry, containing chlorine or dialkylamido groups, or with a free coordination site in the apical position, have been synthesised by salt metathesis and amine elimination. Products with threefold symmetry were generally obtained for tetravalent titanium, whereas for the aluminium complexes either asymmetric structures with two of the three podand arms taking part in coordination to the metal or symmetric arrangements possessing the full threefold symmetry were formed depending on the steric properties of the ligands. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2006)

Co-reporter:Carole Foltz, Björn Stecker, Guido Marconi, Stéphane Bellemin-Laponnaz, Hubert Wadepohl and Lutz H. Gade  
Chemical Communications 2005 (Issue 40) pp:5115-5117
Publication Date(Web):16 Sep 2005
DOI:10.1039/B509571A
Chiral C3-symmetric trisoxazolines are highly efficient stereodirecting ligands in enantioselective CuII Lewis acid catalysis which is based on the concept of a stereoelectronic hemilability of the divalent copper; in direct comparison with the analogous bisoxazoline systems they are more efficient in the enantioselective α-amination as well as the enantioselective Mannich reaction of prochiral β-ketoesters.
Co-reporter:Arno Tuchbreiter, Helmut Werner and Lutz H. Gade  
Dalton Transactions 2005 (Issue 8) pp:1394-1402
Publication Date(Web):15 Mar 2005
DOI:10.1039/B501069A
The introduction of phenyl groups at different points on carbosilane dendrimers allows their acidolytic conversion to highly reactive triflato groups which in turn are readily substituted by anionic nucleophiles. Core phenylated first–fourth generation dendrimers were synthesized from tri(allyl)phenylsilane by an alternating sequence of hydrosilylation and allylation steps. Similarly, carbosilane dendrimers containing phenyl-Si groups at the branching points and in the periphery were prepared from tetraallylsilane which was hydrosilylated with PhHSiCl2. Reaction of the phenylated dendrimers with triflic acid in toluene cleanly gave the silyl triflate derivatives, provided that the correct stoichiometry of the reagents was used. In the presence of a large excess of triflic acid the SiMe3-end groups are slowly converted to SiMe2(OTf)-units. The proof of concept was provided by the fixation of a {Ph2PCH2} group using the lithiated diphenylphosphinomethanide Ph2PCH2Li, obtained by cleavage of Ph3SnCH2PPh2 with PhLi, as well as a lithiated ether-alcohol functionalized triphos derivative to the core of a third generation carbosilane dendrimer.
Co-reporter:Markus B. Meder, Isabelle Haller and Lutz H. Gade  
Dalton Transactions 2005 (Issue 8) pp:1403-1415
Publication Date(Web):15 Mar 2005
DOI:10.1039/B501070E
Bis(2-pyridylimino)isoindolato (BPI) ligands, containing an alkynyl linker unit which allows their fixation to carbosilane dendrimers and dendrons, were synthesized by reaction of 4-nitrophthalodinitrile with 4-butynol giving the phthalodinitrile derivative 1 containing the linker. These were subsequently reacted with two molar equivalents of 2-amino-4-methylpyridine and 2-amino-4-tbutylpyridine yielding the respective BPI protioligands 2a and 2b. Lithiation with LDA and reaction with Si–Cl or Si–OTf (OTf = triflate) end groups in core or peripheral positions of dendritic carbosilanes gave the endodendrally and expdendrally functionalized dendrimers. Among these the first and second generation dendrimers [G-1]8-exo-4-[CCCH2CH2O]-10-MeBPI (8), [G-1]12-exo-4-[CCCH2CH2O]-10-MeBPI (9) and [G-2]16-exo-4-[CCCH2CH2O]-10-MeBPI (10) were synthesized and fully characterized. The functional dendrimers were metallated by reaction with [(PhCN)2PdCl2] in dichloromethane to give the corresponding pallada-dendrimers.
Co-reporter:Nathanaëlle Schneider, Vincent César, Stéphane Bellemin-Laponnaz, Lutz H. Gade
Journal of Organometallic Chemistry 2005 Volume 690(24–25) pp:5556-5561
Publication Date(Web):1 December 2005
DOI:10.1016/j.jorganchem.2005.06.044
Reaction of a series of directly connected oxazoline–imidazolium salts with potassium tert-butoxide and in the presence of CuBr · SMe2 at −78 °C cleanly gave the corresponding 2-oxazolinyl-(N-mesityl)imidazolidenecopper(I) complexes which are monomeric in solution but aggregate in the solid state. X-ray diffraction studies established a dimeric structure for [{2-(4,4-dimethyl)-oxazolinyl-(N-mesityl)imidazolidene}(bromo)copper(I)]2 (2a) whereas the chiral derivative [{2-(4-S-isopropyl)-oxazolinyl-(N-mesityl)imidazolidene}(bromo)-copper(I)]∝ (2b) forms infinite chains of a coordination polymer.Reaction of a series of directly connected oxazoline–imidazolium salts with potassium tert-butoxide and in the presence of CuBr · SMe2 at −78 °C cleanly gave the corresponding 2-oxazolinyl-(N-mesityl)imidazolidenecopper(I) complexes which are monomeric in solution but aggregate in the solid state.
Co-reporter:Vincent César Dr.;Stéphane Bellemin-Laponnaz Dr.;Hubert Wadepohl
Chemistry - A European Journal 2005 Volume 11(Issue 9) pp:
Publication Date(Web):3 MAR 2005
DOI:10.1002/chem.200500132

The direct coupling of oxazolines and N-heterocyclic carbenes leads to chelating C,N ancillary ligands for asymmetric catalysis that combine both an “anchor” unit and a stereodirecting element. Reacting various N-substituted imidazoles with 2-bromo-4(S)-tert-butyl- and 2-bromo-4(S)-isopropyloxazoline gave the imidazolium precursors of the stereodirecting ancillary ligands. A library of ten different ligand precursors was obtained by using this simple procedure (65–97 % yield). These protioligands were metalated in a subsequent step by reaction with [{Rh(μ-OtBu)(nbd)}2] (nbd=norbornadiene), generated in situ from KOtBu and [{RhCl(nbd)}2] giving the corresponding N-heterocyclic carbene complexes [RhBr(nbd)(oxazolinyl-carbene)] 4 aj in good yields. X-ray diffraction studies of two of the rhodium complexes, 4 d and 4 j, established a distorted square-pyramidal coordination geometry with the bromo ligand occupying the apical position. The rhodium–carbene bond length was found to be 2.070(4) Å (4 d) and 2.012(3) Å (4 j). Complexes 4 aj were treated with AgBF4 in dichloromethane, giving the active cationic square-planar catalysts for the hydrosilylation of ketones. As a reference reaction for the catalyst optimisation, the hydrosilylation of acetophenone with diphenylsilane was studied and the system optimised with respect to the counterion (BF4), solvent (THF) and the silane reducing agent (diphenylsilane). The reaction product (1-phenylethanol) was obtained with the highest enantiomeric excess (ee) by carrying out the reaction at −60 °C, whilst the enantioselectivity drops upon going both to lower and higher temperatures. The observation that the temperature dependence of the ee values goes through a maximum indicated a change in the rate-determining step as the temperature is varied. The determination of the initial reaction rate in the hydrosilylation of acetophenone upon varying the catalyst (4 d) and substrate concentrations at −55 °C established a rate law for the initial conversion which is first-order in both substrates as well as the catalyst (Vi=k[4][PhCOMe][Ph2SiH2]). The catalytic system derived from complex 4 d was found to afford high yields and good enantioselectivities in the reduction of various aryl alkyl ketones (acetophenone: 92 % isolated yield and 90 % ee, 2-naphtyl methyl ketone: 99 % yield, 91 % ee). The selectivity for the reduction of prochiral dialkyl ketones is comparable or even superior to the best previously reported for prochiral nonaromatic ketones; whereas cyclopropyl methyl ketone is hydrosilylated with an enantioselectivity of 81 % ee, the increase of the steric demand of one of the alkyl groups leads to improved ee's, reaching 95 % ee in the case of tert-butyl methyl ketone. Linear chain n-alkyl methyl ketones, which are particularly challenging substrates, are reduced in good asymmetric induction, such as 2-octanone (79 % ee) and even 2-butanone (65 % ee).

Co-reporter:Meike Stöhr Dr.;Markus Wahl;Christian H. Galka Dr.;Till Riehm;Thomas A. Jung Dr. Dr.
Angewandte Chemie 2005 Volume 117(Issue 45) pp:
Publication Date(Web):17 OCT 2005
DOI:10.1002/ange.200502316

Verdünnungsprinzip in zwei Dimensionen: Ein Baustein liefert drei unterschiedliche Aggregationsformen, je nach der Oberflächenkonzentration der molekularen Vorstufe (siehe Bild; ML=Monolagenbedeckung). Durch Tempern bei 300 °C werden sehr robuste 2D-Aggregate auf Metalloberflächen erhalten, die zur Konstruktion hierarchischer Systeme höherer Komplexität verwendet werden können.

Co-reporter:Meike Stöhr, Markus Wahl, Christian H. Galka, Till Riehm, Thomas A. Jung,Lutz H. Gade
Angewandte Chemie International Edition 2005 44(45) pp:7394-7398
Publication Date(Web):
DOI:10.1002/anie.200502316
Co-reporter:Vincent César, Stéphane Bellemin-Laponnaz and Lutz H. Gade  
Chemical Society Reviews 2004 vol. 33(Issue 9) pp:619-636
Publication Date(Web):03 Nov 2004
DOI:10.1039/B406802P
In recent years, N-heterocyclic carbenes (NHC) have proved to be a versatile class of spectator ligands in homogeneous catalysis. Being robust anchoring functions for late transition metals, their ligand donor capacity and their molecular shape is readily modified by variation of the substituents at the N-atoms and the structure of the cyclic backbone. After the first attempts to use chiral NHC ligands in asymmetric catalysis in the late 1990's, which initially met with limited success, several novel structural concepts have emerged during the past two years which have led literally to an explosion of the field. With a significant number of highly selective chiral catalysts based on chiral NHCs having been reported very recently, several general trends in the design of new NHC-containing molecular catalysts for stereoselective transformations in organic synthesis emerge.
Co-reporter:Benjamin D. Ward, Aline Maisse-François, Philip Mountford and Lutz H. Gade  
Chemical Communications 2004 (Issue 6) pp:704-705
Publication Date(Web):13 Feb 2004
DOI:10.1039/B316383K
Reaction of the imidotitanium complexes [Ti(NtBu)(N2Npy)(py)] (1) and [Ti(N-2,6-C6H3iPr2)(N2Npy)(py)] (2) with phenyl acetylene and tolyl acetylene in toluene gave the corresponding {2+2} cycloaddition products [Ti(N2Npy){κ2-N(tBu)CHCR}] (R = Ph: 3, Tol: 4) and [Ti(N2Npy){κ2-N(2,6-C6H3iPr2)CHCR}] (R = Ph: 5, Tol: 6). Complex 6 is the first example of a key intermediate in the anti-Markovnikov addition of a primary amine to a terminal acetylene which has been structurally characterized by X-ray diffraction.
Co-reporter:Markus B. Meder;Lutz H. Gade
European Journal of Inorganic Chemistry 2004 Volume 2004(Issue 13) pp:
Publication Date(Web):5 MAY 2004
DOI:10.1002/ejic.200400012

The copper complexes [Cu(4-MeBPI)(OAc)] (4), [Cu(4-Me-10-tBuBPI)(OAc)] (5) and [Cu(BTI)(OAc)] (6) [BPI = 1,3-bis(2-pyridylimino)isoindole, BTI = 1,3-bis(2-thiazolylimino)isoindole] were prepared by reaction of the protio ligands with copper(II) acetate. Compounds 4 and 6 were characterized by X-ray diffraction, establishing distorted square-planar coordination geometries of the copper ions. Two monoclinic modifications of 6 (6a and 6b) were found, both crystallizing in the space group P21/c, but possessing different cell parameters. In contrast to 6a, which is monomeric in the crystal, the second monoclinic modification 6b has a more complicated crystal structure, which is composed of both monomeric complex units such as those found in 6a and infinite chains of coordination polymers. The copper atoms in the polymeric chains of 6b display fivefold coordination and a ligand polyhedron that is an intermediate form between a trigonal-bipyramidal and a square-pyramidal geometry. The allylic peroxylation of cyclohexene with tBuOOH (70% aqueous solution) catalyzed by 4 and 6 (0.17 mol %) gave tert-butylperoxy-3-cyclohexene with selectivities of 86% and 80% (based on cyclohexene) and turnover frequencies of 63 h−1 and 18 h−1, respectively. The peroxylation reaction is thought to proceed according to a Haber−Weiss radical chain mechanism. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2004)

Co-reporter:Vincent César;Stéphane Bellemin-Laponnaz;Lutz H. Gade
European Journal of Inorganic Chemistry 2004 Volume 2004(Issue 17) pp:
Publication Date(Web):23 JUN 2004
DOI:10.1002/ejic.200400182

Reaction of [Rh(μ-OtBu)(diene)]2 [diene = 1,5-cyclooctadiene (COD) or bicyclo[2.2.1]hepta-2,5-diene (NBD)] with the oxazolinylcarbene ligand precursor 2-(4,4-dimethyl)oxazolinylimidazolium bromide (“Me2carboxH+Br”, 2) yielded the five-coordinate complexes [RhBr(Me2carbox)(diene)] [diene = COD (3), NBD (4)]. Single-crystal X-ray structure analyses established a distorted trigonal bipyramidal configuration for 3, whereas the molecular structure of 4 is distorted square pyramidal. The dynamic properties of these five-coordinate complexes have been investigated by variable temperature 1H NMR spectroscopy, indicating polytopal rearrangements of the five-coordinate complexes. The two compounds were converted into the square-planar cationic complexes [Rh(Me2carbox)(diene)]+ (5 and 6, respectively), which have been isolated and fully characterised. The rhodium complex [RhBr(Me2carbox)(CO)] (7) was obtained in high yield in a one-step reaction of the ligand precursor 2 and [Rh(acac)(CO)2] (acac = acetylacetonate). (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2004)

Co-reporter:Bettina Siggelkow;Markus B. Meder;Christian H. Galka;Lutz H. Gade
European Journal of Inorganic Chemistry 2004 Volume 2004(Issue 17) pp:
Publication Date(Web):23 JUN 2004
DOI:10.1002/ejic.200400179

Reaction of the bis(2-pyridylimino)isoindole derivatives (10-Me)-BPI (1a), (11-Me)-BPI (1b), (11-Br)-BPI (1c), (4-Me)-BPI (1d) and 4-Me-10-tBuBPI (1e) with [PdCl2(PhCN)2] and triethylamine in benzene gave the square-planar palladium(II) complexes [PdCl{(10-Me)-BPI}] (2a), [PdCl{(11-Me)-BPI}] (2b), [PdCl{(11-Br)-BPI}] (2c), [PdCl(4-MeBPI)] (2d) and [PdCl(4-Me-10-tBuBPI)] (2e), respectively. Extraction of the crude product 2b with aqueous sodium carbonate solution led to the formation of the dinuclear carbonato-bridged complex [{(11-Me-BPI)Pd}2(μ-CO3)] (3) which was characterized by an X-ray structure analysis. Reaction of 11-Br-BPI (1c) with a large excess (6 equiv.) of the acetylenes Me3SiCCH, Ph3SiCCH and PhCCH under Sonogashira conditions gave the alkynylated derivatives 11-(Me3SiCC)-BPI (4a), 11-(Ph3SiCC)-BPI (4b) and 11-(PhCC)-BPI (4c), which were metallated with bis(benzonitrile)dichloropalladium(II) to yield the PdII complexes [PdCl{11-(Me3SiCC)-BPI}] (5a), [PdCl{11-(Ph3SiCC)-BPI}] (5b) and [PdCl{11-(PhCC)-BPI}] (5c), respectively. The activity of 2b in the catalytic hydrogenation of C=C double bonds was tested for the reaction with styrene, 1-octene and cyclohexene. The stability of the palladium complex, the reproducibility of the reaction kinetics, the different behaviour towards the three olefins chosen as substrates, as well as the possibility of isolating the non-decomposed catalyst after several catalytic runs, provides circumstantial evidence for molecular catalysis with the BPI-palladium complexes. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2004)

Co-reporter:Clémence Dro;Stéphane Bellemin-Laponnaz Dr.;Richard Welter
Angewandte Chemie 2004 Volume 116(Issue 34) pp:
Publication Date(Web):25 AUG 2004
DOI:10.1002/ange.200460187

Der Aktivität von Enzymen mit Tris(histidin)zink-Zentren eifern chirale Trisoxazolinzink-Zweikernkomplexe nach. Das abgebildete Triflato-Kation ist einer dieser ersten nichtenzymatischen Katalysatoren für die asymmetrische Umesterung von aktivierten Estern. Erste Ergebnisse für die kinetische Racematspaltung von Aminosäurephenylestern belegen die Stereoselektivität der Reaktion (F rosa, N blau, O rot, S gelb, Zn grün).

Co-reporter:Clémence Dro;Stéphane Bellemin-Laponnaz Dr.;Richard Welter
Angewandte Chemie International Edition 2004 Volume 43(Issue 34) pp:
Publication Date(Web):25 AUG 2004
DOI:10.1002/anie.200460187

Mimicking the activity of tris(histidine)zinc-based enzymes, chiral dinuclear trisoxazolinezinc complexes such as the triflato cation (see picture) are the first non-enzymatic catalysts for the asymmetric transesterification of activated esters. First results of the kinetic resolution of amino acid phenolates establish the stereoselectivity for this reaction. (F pink, N blue, O red, S yellow, Zn cyan)

Co-reporter:Callum G. M. Benson, Alex J. Plajer, Raúl García-Rodríguez, Andrew D. Bond, Sanjay Singh, Lutz H. Gade and Dominic S. Wright
Chemical Communications 2016 - vol. 52(Issue 62) pp:NaN9686-9686
Publication Date(Web):2016/07/06
DOI:10.1039/C6CC04805F
Deprotonation of the thialdiphosphazane [SPH(μ-NtBu)]2 with a range of metal-bases gives the stable dianion [S–P(μ-NtBu)]22−, which is valence-isoelectronic with the widely-used [RN-P(μ-NR)]22− ligand. Structural studies show that the new ligand has adaptable hard–soft character with respect to the coordinated metal centre and that its multidentate nature can be exploited to construct large cage architectures.
Co-reporter:Lena Hahn, Simin Öz, Hubert Wadepohl and Lutz H. Gade
Chemical Communications 2014 - vol. 50(Issue 38) pp:NaN4943-4943
Publication Date(Web):2014/03/10
DOI:10.1039/C4CC01254B
Water-soluble tetraazaperopyrene (TAPP) derivatives have been synthesized, which show both high photostability and high fluorescence quantum yields (>80%) in water. Furthermore delivery of the dye into cells demonstrated selective staining of the nucleus.
Co-reporter:Thorsten Gehrmann, Matthias Kruck, Hubert Wadepohl and Lutz H. Gade
Chemical Communications 2012 - vol. 48(Issue 18) pp:NaN2399-2399
Publication Date(Web):2012/01/25
DOI:10.1039/C2CC17594K
Reaction of a zirconium imido and a hydrazinediido complex with bis(trimethylsilyl)sulfur diimide yielded the corresponding formal [2+2] cycloaddition products, the trisimidosulfito complex and the hydrazidobis(imido)sulfito complex containing an unprecedented SN2(N–N) unit.
Co-reporter:Markus Wahl, Meike Stöhr, Hannes Spillmann, Thomas A. Jung and Lutz H. Gade
Chemical Communications 2007(Issue 13) pp:NaN1351-1351
Publication Date(Web):2007/02/27
DOI:10.1039/B700909G
Fourfold symmetric zinc-octaethylporphyrin (OEP) has been incorporated in the holes of the hexagonal molecular network generated by thermal dehydrogenation of 4,9-diaminoperylene-quinone-3,10-diimine (DPDI) on a Cu(111) surface and displayed hindered rotation; the reorganization between the potential minima, a rotation–libration, which is characterized by an activation energy of ED = 0.17 ± 0.03 eV, has been monitored in the STM tunnelling currents as a bi-state “switching”.
Co-reporter:Christoph A. Rettenmeier, Hubert Wadepohl and Lutz H. Gade
Chemical Science (2010-Present) 2016 - vol. 7(Issue 6) pp:NaN3542-3542
Publication Date(Web):2016/02/11
DOI:10.1039/C5SC04644K
The study is aimed at a deeper understanding of the electronic structure of the T-shaped nickel(I) complex [LigiPr(iso)Ni] (1b), bearing the iso-PyrrMeBox (bis(oxazolinylmethylidene)pyrrolidinido) pincer ligand, and its CO adduct [LigiPr(iso)Ni(CO)] (2b) as well as to provide insight into the mechanism of autoxidation of the different nickel peroxo species of this ligand type. CO was found to react reversibly with complex 1b resulting in the corresponding CO adduct 2b. The EPR data as well as the results of DFT modeling revealed significant differences in the electronic structure of 1b and 2b. Reaction of [LigPh(iso)Ni] and [LigiPr(iso)Ni] (1a and b) with dioxygen yielded the 1,2-μ-peroxo complexes [Lig(iso)NiO]23a and b which reacted with hydrogen peroxide to give the hydroperoxo complexes [Lig(iso)NiOOH] 5a and b. Thermal aerobic decomposition of the peroxo species 3a and 5a in the presence of O2 led to a C–H activation of the ligand at the benzylic position of the oxazoline ring forming diastereomeric cyclic peroxo complexes 6 and 6′. For the 1,2-μ-peroxo complex 3b the autoxidation of the pincer in the absence of O2 occurred at the tertiary C–H bond of the iPr-group and led to a selective formation of the terminal hydroxo complex [LigiPr(iso)NiOH] 7b and the cyclic alkoxy complex 8 in equimolar quantities, while the corresponding cyclic peroxo species 9 was formed along with 7b in the presence of oxygen. Whether or not O–O bond cleavage occurred in the generation of 9 was established upon performing labeling experiments which indicate that the transformation does not involve an initial O–O bond cleaving step. Based on these observations and a series of stoichiometric transformations a tentative proposal for the processes involved in the anaerobic and aerobic decomposition of 3b has been put forward. Finally, the nickel(II) methyl complex [LigPh(iso)NiMe] 14 reacted with O2 to give the methylperoxo complex [LigPh(iso)NiOOMe] 15 which slowly converted to a mixture of near equal amounts of the formato and the hydroxo complexes, [LigPh(iso)NiOOCH] 16 and [LigPh(iso)NiOH] 7a, along with half an equivalent of methanol. The formato complex 16 itself decomposed at elevated temperatures to CO2, dihydrogen as well as the nickel(I) species 1a.
Co-reporter:Till Riehm, Gabriele De Paoli, Hubert Wadepohl, Luisa De Cola and Lutz H. Gade
Chemical Communications 2008(Issue 42) pp:NaN5350-5350
Publication Date(Web):2008/09/19
DOI:10.1039/B811864G
A new class of N–B–N functionalized perylenes derived from N,N′,N″,N‴-diborylene-3,4,9,10-tetraaminoperylene (DIBOTAP) is readily accessible via several synthetic routes; they display intense green fluorescence and undergo two electrochemically reversible one-electron oxidation steps at ∼−0.2 V and 0.3 V (vs.SCE).
Co-reporter:Nora Grüger, Hubert Wadepohl and Lutz H. Gade
Dalton Transactions 2012 - vol. 41(Issue 46) pp:NaN14030-14030
Publication Date(Web):2012/10/12
DOI:10.1039/C2DT32199H
The novel chelating PNP pincer ligand 2,5-bis((diphenylphosphino)methyl)-1H-pyrrole (1) was prepared and its nickel coordination chemistry explored. The reaction with Ni(COD)2 led to a diamagnetic dinuclear nickel(I) complex (4) which was also obtained by the reaction of the square planar NiII complexes [(PNP-Ph2)NiX] (X = Cl (2), X = I (3)) with Li(Et3BH).
Co-reporter:Matthias Kruck, Désirée C. Sauer, Markus Enders, Hubert Wadepohl and Lutz H. Gade
Dalton Transactions 2011 - vol. 40(Issue 40) pp:NaN10415-10415
Publication Date(Web):2011/06/20
DOI:10.1039/C1DT10617A
Condensation of phthalodinitrile and 2-amino-5,6,7,8-tetrahydroquinoline gave the bis(2-pyridylimino)isoindole protioligand 1 (thqbpiH) in high yield. Deprotonation of thqbpiH (1) using LDA in THF at −78 °C yields the corresponding lithium complex [Li(THF)(thqbpi)] (2) in which the lithium atom enforces almost planar arrangement of the tridentate ligand, with an additional molecule of THF coordinated to Li. Reaction of cobalt(II) chloride or iron(II) chloride with one equivalent of the lithium complex 2 in THF led to formation of the metal complexes [CoCl(THF)(thqbpi)] (3a) and [FeCl(THF)(thqbpi)] (3b). The paramagnetic susceptibility of 3a,b in solution was measured by the Evans method (3a: μeff = 4.17 μB; 3b: μeff = 5.57 μB). Stirring a solution of 1 and cobalt(II) acetate tetrahydrate in methanol yielded the cobalt(II) complex 4 which was also accessible by treatment of 3a with one equivalent of silver or thallium acetate in DMSO. Whereas 3a,b were found to be mononuclear in the solid state, the acetate complex 4 was found to be dinuclear, the two metal centres being linked by an almost symmetrically bridging acetate. For all transition metal complexes paramagnetic 1H as well as 13C NMR spectra were recorded at variable temperatures. The complete assignment of the paramagnetic NMR spectra was achieved by computation of the spin densities within the complexes using DFT. The proton NMR spectra of 3a and 3b displayed dynamic behaviour. This was attributed to the exchange of coordinating solvent molecules by an associative mechanism which was analysed using lineshape analysis (ΔS≠= −154 ± 25 J mol−1 K−1 for 3a and ΔS≠ = −168 ± 15 J mol−1 K−1 for 3b).
Co-reporter:Katharina Weitershaus, Benjamin D. Ward, Raphael Kubiak, Carsten Müller, Hubert Wadepohl, Sven Doye and Lutz H. Gade
Dalton Transactions 2009(Issue 23) pp:NaN4602-4602
Publication Date(Web):2009/04/15
DOI:10.1039/B902038A
A series of new titanium half sandwich complexes, containing a 2-aminopyrrolinato ligand {NXylN}− as the ancillary ligand, have been prepared and are shown to be pre-catalysts for the hydroamination of alkynes. The coordination of {NXylN}− to titanium was achieved by reaction of [Cp*TiMe3] with the protioligand NXylNH giving [Cp*Ti(NXylN)(Me)2] (1). Upon reaction of complex 1 with an excess of tert-butylamine, the imido complex [Cp*Ti(NXylN)(NtBu)(NH2tBu)] (2) was formed. The latter provided the preparative entry to the synthesis of a range of N-aryl substituted imido complexes. Imido ligand exchange with 2,6-dimethylaniline, 2,4,6-trimethylaniline as well as 2,6-diisopropylaniline gave the corresponding arylimido complexes 3–5 in clean reactions. Reaction of the titanium imido complex [Cp*Ti(NXylN)(NtBu)(NH2tBu)] 2 with terminal arylacetylenes, such as phenylacetylene and tolylacetylene, led to C–H activation and the formation of alkynyl/amido complexes, whereas the arylimido complexes 3 and 5 cleanly underwent {2 + 2} cycloaddition, giving the azatitanacyclobutene derivatives. A single-crystal X-ray structure analysis of the azatitanacyclobutene [Cp*Ti(NXylN){κ2N(2,6-C6H3Me2)CTolCH}] (11) provided the first crystallographically characterized Markovnikov cycloaddition product of an imidotitanium complex with a terminal alkyne. The mechanistic aspects of the hydromanination of alkynes with the new Ti half sandwich complexes were studied and established a reversible {2 + 2} cycloaddition step and the cleavage of the metallacyclic intermediate as the rate determining step in the catalytic cycle. The titanium half sandwich imido complexes were found to be active catalysts for the inter- and intramolecular hydroamination of a broad range of alkynes and ω-aminoalkynes.
Co-reporter:Heike Herrmann, Hubert Wadepohl and Lutz H. Gade
Dalton Transactions 2008(Issue 16) pp:NaN2119-2119
Publication Date(Web):2008/03/13
DOI:10.1039/B800682B
The structures and properties of a series of new zirconium hydrazido(1−) complexes and the possibility of converting them to the respective hydrazido(2−) species are reported. Reaction of complex [Zr(N2TBSNpy)Cl2] (1) with the monolithiated hydrazide LiNHNMe2 gave the hydrazido(1−) complex [Zr(N2TBSNpy)(NHNMe2)Cl] (2) which exists as two isomeric forms (2a and 2b) in solution. All attempts to convert a mixture of 2a and 2b to the respective hydrazido(2−) compound by reaction with the bulky base lithium hexamethyldisilazide or via the alkyl/hydrazido(1−) complex [Zr(N2TBSNpy)(CH2SiMe3)(NHNMe2)] (3) and subsequent thermal alkane elimination failed. Reaction of 1 with LiHNNPhMe gave a mixture of stereoisomers of [Zr(N2TBSNpy)(NHNMePh)Cl] (4a and 4b), in which the hydrazido unit is end-on bound in solution and η2-bonded in the solid state. Reaction of this mixture with lithium hexamethyldisilazide in the presence of pyridine selectively yielded the hydrazido(2−) complex [Zr(N2TBSNpy)(NNPhMe)(py)] (5) which aggregated upon attempts to isolate it. Reaction of the insoluble precipitate with 4-dimethylaminopyridine (dmap) selectively gave the corresponding hydrazido(2−) complex [Zr(N2TBSNpy)(NNPhMe)(dmap)] (6), which could be obtained in a one-pot reaction directly from 1 and which was analytically and spectroscopically fully characterized. It appears that the isolation of stable hydrazido(2−) complexes of zirconium depends on the type of substituents at the Nβ atom as well as the co-ligands coordinated to the metal centre.
Co-reporter:Heike Herrmann, Thorsten Gehrmann, Hubert Wadepohl and Lutz H. Gade
Dalton Transactions 2008(Issue 44) pp:NaN6241-6241
Publication Date(Web):2008/09/30
DOI:10.1039/B807808D
Reaction of the diamidozirconium complex [Zr(N2TBSNpy)(NMe2)2] (1) (N2TBSNpy = CH3C(C5H4N)(CH2NSiMe2tBu)2) or the diamidohafnium complex [Hf(N2TBSNpy)(NMe2)2] (2) with one molar equiv. of 1-aminopyridinium triflate in the presence of one equiv. of pyridine gave the corresponding (1-pyridinio)imido complexes [Zr(N2TBSNpy)(N–NC5H5)(OTf)(py)] (3) and [Hf(N2TBSNpy)(N–NC5H5)(OTf)(py)] (4). These were converted to the acetylide complexes [Zr(N2TBSNpy)(N–NC5H5)(CCPh)(py)] (5) and [Hf(N2TBSNpy)(N–NC5H5)(CCPh)(py)] (6) by reaction with lithium phenylacetylide and substitution of the triflato ligand. Upon reaction of 3 and 4 with one molar equivalent of R-NC (R = tBu, Cy, 2,6-xyl), N–N bond cleavage in the (1-pyridinio)imido unit took place and the respective carbodiimido complexes [M(N2TBSNpy](NCNR)(OTf)(py)] (7–12) were formed instantaneously. A similar type of reaction with CO gave the isocyanato complex [Zr(N2TBSNpy](NCO)(OTf)(py)] (13). Finally, the abstraction of the pyridine ligand in compounds 3 and 4 with B(C6F5)3 led to the formation of the triflato-bridged dinuclear complexes [Zr(N2TBSNpy)(N–NC5H5)(OTf)]2 (14) and [Hf(N2TBSNpy)(N–NC5H5)(OTf)]2 (15).
Co-reporter:Lenka Lukešová, Benjamin D. Ward, Stéphane Bellemin-Laponnaz, Hubert Wadepohl and Lutz H. Gade
Dalton Transactions 2007(Issue 9) pp:NaN922-922
Publication Date(Web):2007/01/18
DOI:10.1039/B700269F
The thulium complexes [Tm(iPr-trisox)(CH2SiMe2R)3] (R = Me 1a, Ph 1b) were synthesized from the thulium trialkyl precursors [Tm(CH2SiMe2R)3(thf)2]; reaction of 1a with two equivalents of [Ph3C][B(C6F5)4] gave a cationic complex 1c, which was found to polymerize 1-hexene, 1-heptene and 1-octene to give the corresponding polyolefins with moderate to good activities and with minimum isotacticity of 90%, 83% and 95%, respectively.
Co-reporter:Michaela Kilian, Hubert Wadepohl and Lutz H. Gade
Dalton Transactions 2008(Issue 5) pp:NaN584-584
Publication Date(Web):2007/11/14
DOI:10.1039/B716064J
Reaction of the triamido stannate MeSi{SiMe2N[(R)-CHMePh]}3SnLi (1) with 0.5 molar equivalent of [RhCl(olefin)2]2 (olefin = COE, C2H4) or [RhCl(PiPr3)2]2 yielded the Rh–Sn complexes [MeSi{SiMe2N[(R)-CHMePh]}2{SiMe2N[(R)-CHMe(η6-C6H5)}SnRh(L)] (L = COE: 2a, C2H4: 2b, PiPr33); their intramolecular η6-coordination, along with the tin–rhodium bond, represents the first “ansa” π-arene/stannate system.
3-Oxabicyclo[3.1.0]hexan-2-one, 6-phenyl-, (1S,5R,6S)-
2-Quinolinamine,5,6,7,8-tetrahydro-
Cyclohexanepropanenitrile, 2-(hydroxyimino)-
Oxazole, 4,5-dihydro-4-phenyl-, (4R)-
Oxazole, 4,5-dihydro-4-(1-methylethyl)-, (4S)-
Acetic acid, diazo-, (2E)-3-phenyl-2-propenyl ester
(4S,4'S)-2,2'-(Propane-2,2-diyl)bis(4-phenyl-4,5-dihydrooxazole)
Butanedioic acid,2-(phenylmethyl)-, 1,4-dimethyl ester, (2R)-