Co-reporter:Philipp Zimmermann and Christian Limberg
Journal of the American Chemical Society March 29, 2017 Volume 139(Issue 12) pp:4233-4233
Publication Date(Web):February 7, 2017
DOI:10.1021/jacs.6b12434
The activation and selective transformation of virtually inexhaustible or easy-to-generate chemicals like N2, O2, CO2, CO, H2, or methane gas to value-added products is a lively area of current research, because of its economic relevance as well as its huge ecological impact. Biologists and chemists have put forth a lot of effort toward understanding and modeling the mechanisms of biological small-molecule activation, and in several catalytic cycles proposed for nickel-containing enzymes, nickel(I) plays a key role. In recent years also in synthetic chemistry the huge potential of complex nickel(I) units for the activation and transformation of small molecules has been discovered and exploited. This Perspective highlights some representative examples of nickel(I)-based small-molecule activation, intending to establish awareness of the competencies and scope of nickel(I) compounds.
Co-reporter:D. Pinkert;M. Keck;S. Ghassemi Tabrizi;C. Herwig;F. Beckmann;B. Braun-Cula;M. Kaupp;C. Limberg
Chemical Communications 2017 vol. 53(Issue 57) pp:8081-8084
Publication Date(Web):2017/07/13
DOI:10.1039/C7CC04670G
The famous α-Fe active sites in Fe-zeolites have recently been revealed to correspond to mononuclear high-spin iron(II) centres in square planar coordination environments. Here we report a first iron siloxide complex which represents a faithful structural and spectroscopic model of such sites. Notably, also an allogon with a distorted structure exists and could be crystallised.
Co-reporter:S. Hoof;M. Sallmann;C. Herwig;B. Braun-Cula;C. Limberg
Dalton Transactions 2017 vol. 46(Issue 48) pp:16792-16795
Publication Date(Web):2017/12/12
DOI:10.1039/C7DT04056C
To support mechanistic inferences made for an iron-based dioxygenase model, a nickel analogue, i.e. a TpNi-malonate (1) was prepared. 1 proved to represent a rare case of a nickel complex reacting with O2 in a controlled manner – mechanistically different from the iron case – and leads to hydroxylation of the malonate.
Co-reporter:Georg Bendt, Rafael Schiwon, Soma Salamon, Joachim Landers, Ulrich Hagemann, Christian Limberg, Heiko Wende, and Stephan Schulz
Inorganic Chemistry 2016 Volume 55(Issue 15) pp:7542-7549
Publication Date(Web):July 8, 2016
DOI:10.1021/acs.inorgchem.6b00951
Nearly phase-pure bismuth ferrite particles were formed by thermolysis of the single-source precursor [Cp(CO)2FeBi(OAc)2] (1) in octadecene at 245 °C, followed by subsequent calcination at 600 °C for 3 h. In contrast, the slightly modified compound [Cp(CO)2FeBi(O2CtBu)2] (2) yielded only mixtures of different bismuth oxide phases, revealing the distinctive influence of molecular design in material synthesis. The chemical composition, morphology, and crystallinity of the resulting materials were investigated by X-ray diffraction, transmission electron microscopy, and energy-dispersive X-ray spectroscopy. In addition, the optical properties were investigated by Fourier transform infrared and UV–vis spectroscopies, showing a strong band gap absorption in the visible range at 590 nm (2.2 eV). The magnetic behavior was probed by vibrating-sample and superconducting quantum interference device magnetometry, as well as 57Fe Mössbauer spectroscopy.
Co-reporter:Carolin Tschersich, Santina Hoof, Nicolas Frank, Christian Herwig, and Christian Limberg
Inorganic Chemistry 2016 Volume 55(Issue 4) pp:1837-1842
Publication Date(Web):February 4, 2016
DOI:10.1021/acs.inorgchem.5b02740
To investigate the effect of X in ambiphilic compounds BiX(o-PPh2-C6H4)2, PBiP-X, on metallophilic Pt–Bi interactions in its PtCl2 complexes two new derivatives PBiP-Me and PBiP-C6F5 were synthesized. Reaction with dichloro(1,5-cyclooctadiene)platinum(II) led to the platinum(II) complexes [PtCl2(PBiP-Me)], 3, and [PtCl2(PBiP-C6F5)], 4, which together with the halide [PtCl2(PBiP-Cl)], 2, reported previously, establish a series of related PBiP-X complexes differing only in X. This could be complemented by accessing [PtCl2(PBiP-OTf)], 5, through the reaction of 2 with AgOTf. Analysis of the geometrical and electronic structures of these complexes revealed that in all cases the platinum(II) centers act as donors (through their filled dz2 orbitals) to the bismuth(III) centers (possessing σ*(Bi-X)/6p acceptor orbitals). The strength of these interactions increases with increasing electron-withdrawing character of X, which supports the conceptual approach in constructing this new class of compounds.
Co-reporter:Henrike Gehring, Ramona Metzinger, Beatrice Braun, Christian Herwig, Sjoerd Harder, Kallol Ray and Christian Limberg
Dalton Transactions 2016 vol. 45(Issue 7) pp:2989-2996
Publication Date(Web):13 Jan 2016
DOI:10.1039/C5DT04266F
After lithiation of PYR-H2 (PYR = [(NC(Me)C(H)C(Me)NC6H3(iPr)2)2(C5H3N)]2−) – the precursor of an expanded β-diketiminato ligand system with two binding pockets – with KN(TMS)2 the reaction of the resulting potassium salt with FeBr2 led to a dinuclear iron(II) bromide complex [(PYR)Fe(μ-Br)2Fe] (1). Through treatment with KHBEt3 the bromide ligands could be replaced by hydrides to yield [PYR)Fe2(μ-H)2] (2), a distorted analogue of known β-diketiminato iron hydride complexes, as evidenced by NMR, Mößbauer and X-ray absorption spectroscopy, as well as by its reactivity: for instance, 2 reacts with the proton source lutidinium triflate via protonation of the hydride ligands to form an iron(II) product [(PYR)Fe2(OTf)2] (4), while CO2 inserts into the Fe–H bonds generating the formate complex [(PYR)Fe2(μ-HCOO)2] (5); in the presence of traces of water partial hydrolysis occurs so that [(PYR)Fe2(μ-OH)(μ-HCOO)] (6) is isolated. Altogether, the iron(II) chemistry supported by the PYR2− ligand is distinctly different from the one of nickel(II), where both, the arrangement of the two binding pockets and the additional pyridyl donor led to diverging features as compared with the corresponding system based on the parent β-diketiminato ligand.
Co-reporter:Rafael Schiwon, Fabian Schax, Beatrice Braun, Christian Limberg
Journal of Organometallic Chemistry 2016 Volume 821() pp:71-77
Publication Date(Web):15 October 2016
DOI:10.1016/j.jorganchem.2016.02.041
•Synthesis of the heterobimetallic Fe–Bi formate [Fp–Bi(O2CH)2]n.•Crystallization and characterization of two polymorphs of [Fp–Bi(O2CH)2]n.•Formation of a heterotrimetallic iron-bismuth-rhenium compound [Fp–Bi(Cl)ReCp2].•Synthesis of the heterobimetallic Re–Bi acetate [Cp2Re–Bi(O2CCH3)2].The reaction of [Fp]2 (Fp = (η5-C5H5)(CO)2Fe) with bismuth formate was shown to proceed via Fe-Bi bond formation yielding [Fp–Bi(O2CH)2]n, 1, that could be crystallized in form of two polymorphs. These differ in the way the [Fp−Bi(O2CH)2] units are connected via bridging formate ligands to yield the respective coordination polymers. The acetate derivative [Fp–Bi(O2CCH3)2] was used as a precursor compound targeting at the construction of an additional transition metal-bismuth bond: the reaction of [Fp–Bi(O2CCH3)2] with [Cp2ReH] likely yields in the envisaged complex [Fp−Bi(O2CCH3)ReCp2] first but then undergoes further transformation due to the presence of chloride and acetic acid in solution, which lead to the formation of [Fp−Bi(Cl)ReCp2], 3, and [Cp2Re–Bi(O2CCH3)2], 4, respectively. The identity of the heterotrimetallic compound 3 was established by a single-crystal X-ray diffraction analysis, while an independent synthetic route (reaction of [Cp2ReH] with [Bi(O2CCH3)3]) was developed for 4 in order to unambiguously identify 4 also as the product of the reaction between [Fp–Bi(O2CCH3)2] and [Cp2ReH].The iron bismuth formate [Fp–Bi(O2CH2)2)]n was accessed through reaction of Fp2 with bismuth formate and found to crystallize in two polymorphic structures; crystallization procedures to obtain both polymorphs as pure phases were developed. The reaction of the acetate [Fp–Bi(O2CCH3)2] with rhenocene hydride was investigated and led to the formation of the unique heterotrimetallic compound [Fp–Bi(Cl)ReCp2].
Co-reporter:Madleen Sallmann and Christian Limberg
Accounts of Chemical Research 2015 Volume 48(Issue 10) pp:2734
Publication Date(Web):August 25, 2015
DOI:10.1021/acs.accounts.5b00148
Mononuclear, O2-activating nonheme iron enzymes are a fascinating class of metalloproteines, capable of realizing the most different reactions, ranging from C–H activation, via O atom transfer to C–C bond cleavage, in the course of O2 activation. They can lead us the way to achieve similar reactions with comparable efficiency and selectivity in chemical laboratories, which would be highly desirable aiming at accessing value-added products or to achieve degradation of unwanted compounds. Hence, these enyzmes motivate attempts to construct artificial low-molecular weight analogues, mimicking structural or functional characteristics. Such models can, for instance, provide insights about which of the features inherent to an active site are essential and guarantee the enzyme function, and from this kind of information the minimal requirements for a biomimetic or bioinspired complex that may be applied in catalysis can be derived. On the other hand, they can contribute to an understanding of the enzyme functioning. In order to create such replicates, it is important to faithfully mimic the surroundings of the iron centers in their active sites. Most of them feature two histidine residues and one carboxylate donor, while a few exhibit a deceptively simple (His)3Fe active site. For the simulation of these, the trispyrazolyl borate ligand (Tp) particularly offers itself, as the facial arrangement of three pyrazole donors is reminiscent of the three histidine-derived imidazole donors. The focus of this Account will be on bioinorganic/biomimetic research from our laboratory utilizing Tp ligands to develop molecular models for (i) two representatives of the (His)3Fe-enzyme family, namely, the cysteine dioxygenase (CDO) and acetyl acetone dioxygenase (Dke1), (ii) a related but less well-explored variant of the CDO—the 2-aminoethanethiol dioxygenase—as well as (iii) the 2-His-1-carboxylate representative 1-aminocyclopropane-1-carboxylic acid oxidase (ACCO). The CDO catalyzes the dioxygenation of cysteine with O2 to give cysteine sulfinic acid, which could be mimicked at TpFe units in a realistic manner. Furthermore, the successful dioxygenation of 2-aminoethanethiol at the same complex metal fragments lends further support to the hypothesis that the active sites of CDO and the one of 2-aminoethanethiol dioxygenase, whose structure is unknown, are quite similar.Dke1 is capable of cleaving diketones and ketoesters to give the corresponding carboxylic acids and α-keto aldehydes, and Tp-based models have achieved comparable C–C bond cleavage reactions. The ACCO develops ethylene from ACC in the course of oxidation, and recently this has been achieved the first time for a TpFe model, too.
Co-reporter:M. Sallmann, B. Braun and C. Limberg
Chemical Communications 2015 vol. 51(Issue 31) pp:6785-6787
Publication Date(Web):19 Mar 2015
DOI:10.1039/C5CC01083G
A novel complex TpMe,PhFe(SCH2CH2NH2) has been synthesized as a speculative model for ADO. Indeed its reaction with O2 led to the dioxygenation of the S atom and thus to hypotaurine. This finding may allow us to draw conclusions on the constitution of the ADO active site, whose structure is still unknown.
Co-reporter:Carolin Tschersich, Beatrice Braun, Christian Herwig, Christian Limberg
Journal of Organometallic Chemistry 2015 Volume 784() pp:62-68
Publication Date(Web):15 May 2015
DOI:10.1016/j.jorganchem.2014.09.019
•PBiP ligand enables synthesis of heterometallic M–Bi complexes (M = CuI, AgI).•MCl (M = Cu, Ag) as precursor salt leads to dimeric [M(μ-Cl)(PBiP)]2 compounds.•The metallophilic M⋯Bi interactions are weaker for Cu and Ag than for Au and mainly M ← Bi in nature.•Setting out from MX (X = BF4, OTf) the synthesis of monomeric complexes was achieved in solution.•M–Bi monomers crystallize with chain-like structures [M(anion)(PBiP)]x.Reaction of the pincer ligand PBiP with MCl (M = Cu, Ag) led to the complexes [M(μ-Cl)(PBiP)]2 featuring M⋯Bi interactions, which were shown by DFT/NBO to have predominantly Bi → M character. In the series of compounds [Cu(μ-Cl)(PBiP)]2, [Ag(μ-Cl)(PBiP)]2 and [AuCl(PBiP)] the ratios between the M–Bi distances and the covalent/van der Waals radii decrease, which indicates a strengthening of the interactions. Employing metal precursors with weakly interacting anions leads, according to NMR spectroscopic results, to monomeric complexes, which, however, aggregate in course of the crystallization process to give [M(anion)(PBiP)]x.Employing the PBiP pincer ligand for the coordination of Cu+ and Ag+ leads to heterometallic complexes featuring predominantly Bi → Cu(Ag) interactions. While these complexes dimerize via the chloride ligands, reaction with chloride free metal precursors leads to monomeric complexes in solution, which crystallize in chain-like structures aggregated via the bismuth-bound chloride.
Co-reporter:Dipl.-Chem. Fabian Schax;B.Sc. Simon Suhr;Dr. Eckhard Bill;Dr. Beatrice Braun;Dr. Christian Herwig;Dr. Christian Limberg
Angewandte Chemie International Edition 2015 Volume 54( Issue 4) pp:1352-1356
Publication Date(Web):
DOI:10.1002/anie.201409294
Abstract
The reaction of 1,1,3,3-tetraphenyl-1,3-disiloxandiol (LH2) with n-butyllithium and CrCl2 results in a mononuclear chromium(II) complex (1) that further reacts with O2 at low temperatures to yield a mononuclear chromium(III) superoxide complex [L2CrO2(THF)][Li2(THF)3] (2). The crystal structure revealed that the chromium superoxido entity is stabilized by the coordination to an adjacent lithium cation. Complex 2 thus contains an unprecedented heterobimetallic [CrIII(μ-O2)Li+] core; beyond this it is the first chromium superoxide for which a temperature-dependent magnetic characterization could be achieved, and the first structurally characterized representative with chromium in an exclusive O-donor environment.
Co-reporter:Dr. Aline Arnold;Dr. Ramona Metzinger ;Dr. Christian Limberg
Chemistry - A European Journal 2015 Volume 21( Issue 3) pp:1198-1207
Publication Date(Web):
DOI:10.1002/chem.201405155
Abstract
New tripodal ligand L2 featuring three different pyridyl/imidazolyl-based N-donor units at a bridgehead C atom, from which one of the imidazolyl units is separated by a phenylene linker, was synthesized and investigated with regards to copper(I) complexation. The resulting complex [(L2)Cu]OTf (2OTf), the known complex [(L1)Cu]OTf (1OTf; L1 differs from L2 in that it lacks the phenylene spacer) and [(L3)Cu]OTf (3OTf), prepared from a known chiral, tripodal, N-donor ligand featuring pyridyl, pyrazolyl, and imidazolyl donors, were tested as catalysts for the oxidation of sodium 2,4-di-tert-butylphenolate (NaDTBP) with O2. Indeed, they mediated NaDTBP oxidation to give mainly the corresponding catecholate and quinone (Q). None of the complexes 1OTf, 2OTf, and 3OTf is superior to the others, as yields were comparable and, if the presence of protons is guaranteed by concomitant addition of the phenol DTBP, the oxidation can also be performed catalytically. For all complexes stoichiometric oxidations under certain conditions (concentrated solutions, high NaDTBP content) were found to also generate products typical for metal-mediated intradiol cleavage of the catecholate with O2. As shown representatively for 1OTf this dioxygenation sets in at a later stage of the reaction. Initially a copper species responsible for the monooxygenation must form from 1OTf/NaDTBP/O2, and only thereafter is the copper species responsible for dioxygenation formed and consumes Q as substrate. Hence, under these circumstances complexes 1OTf–3OTf show both monooxygenase and catechol dioxygenase activity.
Co-reporter:Dipl.-Chem. Fabian Schax;B.Sc. Simon Suhr;Dr. Eckhard Bill;Dr. Beatrice Braun;Dr. Christian Herwig;Dr. Christian Limberg
Angewandte Chemie 2015 Volume 127( Issue 4) pp:1368-1372
Publication Date(Web):
DOI:10.1002/ange.201409294
Abstract
Durch die Reaktion von 1,1,3,3-Tetraphenyl-1,3-disilanol (LH2) mit n-Butyllithium und CrCl2 bildet sich ein einkerniger CrII-Komplex 1, der bei niedrigen Temperaturen mit O2 zu einem einkernigen CrIII-Superoxid-Komplex [L2CrO2(THF)][Li2THF3] (2) weiterreagiert. Durch Kristallstrukturanalyse konnte gezeigt werden, dass die Chromsuperoxid-Einheit durch die Koordination an ein benachbartes Lithiumkation stabilisiert wird. 2 weist damit eine neuartige heterobimetallische [CrIII(μ-O2)Li]-Einheit auf; des Weiteren handelt es sich bei 2 um das erste Chromsuperoxid, dessen magnetische Eigenschaften in Abhängigkeit von der Temperatur bestimmt werden konnten, und um den ersten strukturell beschriebenen Vertreter, in dem Chrom ausschließlich von O-Donoren umgeben ist.
Co-reporter:Carolin Tschersich, Beatrice Braun, Christian Herwig, and Christian Limberg
Organometallics 2015 Volume 34(Issue 15) pp:3782-3787
Publication Date(Web):July 20, 2015
DOI:10.1021/acs.organomet.5b00439
The compound BiCl(o-PPh2-C6H4)2, PBiP-Cl, which in previous work had been shown to form complexes with pronounced M→Bi character, when metal(I) ions of group 11 or PtII and PdII ions were coordinated, behave differently in contact with late metal atoms in low oxidation states, known to easily undergo oxidative additions: Treatment of PBiP-Cl with M(PPh3)4, with M = Pt, Pd, led to the formal insertion of M into the Bi–Cl bond to yield complexes [MCl(PBiP)]. Analogues PBiP-X with X = Br, I that could be accessed behaved similarly, producing [MX(PBiP)]. Both types of complexation reactions—coordination of PBiP-Cl as an ambiphilic ligand and oxidative addition—were observed to occur when [Ir(acac)(cod)] was chosen as the precursor compound. NMR investigations clearly indicated the presence of [IrI(acac)(PBiP-Cl)] and [IrIII(acac)Cl(PBiP)] beside each other in solution, from which, however, only [IrIII(acac)Cl(PBiP)] could be crystallized. DFT results showed that both products differ only slightly in energy. Reaction of PBiP-Cl with [Ir(acac-F6)(cod)] led only to the iridium(III) product, underlining that electronic effects sensitively influence the course of reactivity and the position of the equilibrium.
Co-reporter:Rafael Schiwon, Katharina Klingan, Holger Dau and Christian Limberg
Chemical Communications 2014 vol. 50(Issue 1) pp:100-102
Publication Date(Web):11 Nov 2013
DOI:10.1039/C3CC46629A
Modification of the Co-oxo cores of cobalt-polyoxometalate water oxidation catalysts is detectable by X-ray absorption spectroscopy (XAS) as demonstrated by comparison of Na10[Co4(H2O)2(PW9O34)2] (1) and Na17[((Co(H2O))Co2PW9O34)2(PW6O26)] (2). XAS reveals the integrity of 1 uncompromised by oxidant-driven water oxidation, which proceeds without formation of catalytic cobalt oxide.
Co-reporter:Jan P. Falkenhagen, Beatrice Braun, Eckhard Bill, Dominik Sattler, and Christian Limberg
Inorganic Chemistry 2014 Volume 53(Issue 14) pp:7294-7308
Publication Date(Web):July 1, 2014
DOI:10.1021/ic500584a
The potential of iron molybdates as catalysts in the Formox process stimulates research on aggregated but molecular iron–molybdenum oxo compounds. In this context, [(Me3TACN)Fe](OTf)2 was reacted with (nBu4N)2[MoO4], which led to an oxo cluster, [[(Me3TACN)Fe][μ-(MoO4-κ3O,O′,O″)]]4 (1, Fe4Mo4) with a distorted cubic structure, where the corners are occupied by (Me3TACN)Fe2+ and [Mo═O]4+ units in an alternating fashion, being bridged by oxido ligands. The cyclic voltammogram revealed four reversible oxidation waves that are assigned to four consecutive FeII → FeIII transfers and motivated attempts to isolate compounds containing the respective cations. Indeed, a salt with a FeII2FeIII2MoVI4 constellation, [Fe4Mo4](TCNQ)2 (2), could be isolated after treatment with TCNQ. The FeIIFeIII3MoVI4 stage could be reached via oxidation with DDQ or 3 equiv of thianthrenium hexafluorophosphate (ThPF6), giving [Fe4Mo4](DDQ)3 (4) or [Fe4Mo4](PF6)3 (5), respectively. The fully oxidized FeIII4MoVI4 state was generated through oxidation with 4 equiv of ThPF6, leading to [Fe4Mo4](PF6)4, which showed a unique behavior: upon storage, one of the [Mo═O]4+ corners inverts, so that the terminal oxido ligand is located in the interior of the cage, leading to the formation of [[(Me3TACN)Fe]4[μ-([MoO4]3[MoO4(MeCN-κN)])-κ3O,O′,O″)](PF6)4 (7). In this form, the compound could no longer be employed to enter the cyclic voltammogram recorded for 1, 3, and 5 from the oxidized side; no discrete redox events were observed. Compounds 1–3 and 7 were characterized structurally and 1, 3, and 7 additionally by SQUID measurements and Mössbauer spectroscopy. The data reveal a high degree of charge delocalization. 16O/18O exchange experiments with labeled water performed with 1 revealed an interesting parallel with the Formox catalyst: water−18O exchanges its label with all of the oxido ligands (bridging and terminal). This property relates to the ion mobility being held responsible for the activity of iron molybdate catalysts compared to neat MoO3 or Fe2O3.
Co-reporter:Bettina Horn, Christian Limberg, Christian Herwig, and Beatrice Braun
Inorganic Chemistry 2014 Volume 53(Issue 13) pp:6867-6874
Publication Date(Web):June 10, 2014
DOI:10.1021/ic500698v
Mononuclear nickel(II) thiolate complexes [LtBuNi(SEt)] (1) and [LtBuNi(aet)] (2, aet = −S(CH2)2NH2) (LtBu = [HC(C(tBu)NC6H3(iPr)2)2]−), supported by a bulky nacnac ligand, were synthesized by treatment of the nickel(II) bromide precursor [LtBuNi(Br)] (I) with the potassium salts of ethanethiol and cysteamine, respectively. The nickel atom in 1 features a planar T-shaped environment, while the Ni ion within 2 shows a distorted square planar coordination geometry, as the aminoethanethiolate (aet) is coordinated as a chelating ligand. In 2 the β-diketiminate ligand binds in a rarely observed κ2C,N coordination mode. Reduction of complex 1 or its benzenethiolate analogue [LtBuNi(SPh)] (II) by KC8 resulted in the formation of dinuclear NiI thiolates (K·OEt2)(K)[LtBuNi(SEt)]2 (3) and (K·OEt2)2[LtBuNi(SPh)]2 (4), respectively. In these compounds [LtBuNi(SR)]− units are held together by potassium cations produced in the reduction process. All compounds mentioned were structurally characterized by single-crystal X-ray crystallography.
Co-reporter:Jan P. Falkenhagen, Christian Limberg, Serhiy Demeshko, Sebastian Horn, Michael Haumann, Beatrice Braun and Stefan Mebs
Dalton Transactions 2014 vol. 43(Issue 2) pp:806-816
Publication Date(Web):23 Oct 2013
DOI:10.1039/C3DT52349G
The reaction between [(TPA)Fe(MeCN)2](OTf)2 and [nBu4N](Cp*MoO3) yields the novel tetranuclear complex [(TPA)Fe(μ-Cp*MoO3)]2(OTf)2, 1, with a rectangular [Mo–O–Fe–O–]2 core containing high-spin iron(II) centres. 1 proved to be an efficient initiator/(pre)catalyst for the autoxidation of cis-cyclooctene with O2 to give cyclooctene epoxide. To test, which features of 1 are essential in this regard, analogues with zinc(II) and cobalt(II) central atoms, namely [(TPA)Zn(Cp*MoO3)](OTf), 3, and [(TPA)Co(Cp*MoO3)](OTf), 4, were prepared, which proved to be inactive. The precursor compounds of 1, [(TPA)Fe(MeCN)2](OTf)2 and [nBu4N](Cp*MoO3) as well as Cp2*Mo2O5, were found to be inactive, too. Reactivity studies in the absence of cyclooctene revealed that 1 reacts both with O2 and PhIO via loss of the Cp* ligands to give the triflate salt 2 of the known cation [((TPA)Fe)2(μ-O)(μ-MoO4)]2+. The cobalt analogue 4 reacts with O2 in a different way yielding [((TPA)Co)2(μ-Mo2O8)](OTf)2, 5, featuring a Mo2O84− structural unit which is novel in coordination chemistry. The compound [(TPA)Fe(μ-MoO4)]2, 6, being related to 1, but lacking Cp* ligands failed to trigger autoxidation of cyclooctene. However, initiation of autoxidation by Cp* radicals was excluded via experiments including thermal dissociation of Cp2*.
Co-reporter:Claudia Köthe;Beatrice Braun;Christian Herwig
European Journal of Inorganic Chemistry 2014 Volume 2014( Issue 31) pp:5296-5303
Publication Date(Web):
DOI:10.1002/ejic.201402812
Abstract
The β-diketiminato nickel(II) hydrazido(–1) complex [LtBuNi(η2-N2H3)], (IV) was investigated with respect to its deprotonation/protonation behavior. Deprotonation with KOtBu led to [LtBuNi(μ,η2:η2-N2H2)K(solv)] (1), and this process was proven to be reversible. Spectroscopic and DFT studies revealed an electronic structure intermediate between nickel(II) hydrazido(–2) and nickel(0) diazene. On the other hand, protonation of IV with [LutH]OTf reversibly generated the hydrazine complex [LtBuNi(η2-N2H4)]OTf (2). Warming a solution of IV led to N–N bond cleavage yielding the nickel(I) ammine complex [LtBuNi(NH3)], (3). Hence, LtBuNi moieties were shown to effectively activate NxHy species for diverse conversions.
Co-reporter:Fabian Schax;Beatrice Braun
European Journal of Inorganic Chemistry 2014 Volume 2014( Issue 12) pp:2124-2130
Publication Date(Web):
DOI:10.1002/ejic.201400088
Abstract
A three-step synthetic route for a new tripodal branched trisilanol ligand PhSi(OSiPh2OH)3 (LH3) was developed. X-ray diffraction analysis revealed that the trisilanol crystallizes as a dimer with a cyclic hydrogen-bonding network. The reaction of LH3 with three equivalents of CunMesn (Mes = mesityl) led to a hexanuclear compound [L2Cu6] (1), which was characterized by single-crystal X-ray diffraction analysis as well as by solution NMR spectroscopy. The crystal structure of 1 revealed that the compound features a hexagonal planar CuI6 ring, which is the first of its kind in an oxygen environment. Deprotonation of the ligand with n-butyllithium and subsequent reaction with CuBr2 resulted in the dinuclear CuII complex [L′2Cu2][Li(THF2)]2 (2, THF = tetrahydrofuran), which contains a new siloxide ligand formed from L3– by the elimination of a SiOPh2 unit, as evidenced by X-ray diffraction analysis. To check if the “Ph2SiO” elimination is a general behavior of this trisilanol, the reaction with ZnBr2 was investigated under analogous conditions. However, this led to the isolation of [L2Zn2][Li(OEt)]2 (3) without any rearrangement of the siloxide ligand.
Co-reporter:Denise Pinkert ;Dr. Christian Limberg
Chemistry - A European Journal 2014 Volume 20( Issue 30) pp:9166-9175
Publication Date(Web):
DOI:10.1002/chem.201402826
Abstract
Iron centres incorporated in silicate frameworks or located in their pores have been shown to possess unique catalytic properties. As compared to aluminosilicates this area of zeolite chemistry is much younger and in the first part of this review the findings made so far are summarised. Molecular model compounds may help to understand the formation, corrosion and reactivity of such materials or to even develop new ones. Hence, the subsequent parts deal with molecular iron siloxides, the number of which is still quite limited, and their behaviour also in relation to the iron-modified zeolites is outlined. At first, compounds based on an incompletely condensed cubic silsequioxane are discussed, before iron(III) complexes of more basic siloxide ligands with varying steric demands are described. Finally, recent developments based on branched polydentate siloxides are presented.
Co-reporter:Dr. Peter Haack;Dr. Christian Limberg
Angewandte Chemie International Edition 2014 Volume 53( Issue 17) pp:4282-4293
Publication Date(Web):
DOI:10.1002/anie.201309505
Abstract
Research on O2 activation at ligated CuI is fueled by its biological relevance and the quest for efficient oxidation catalysts. A rarely observed reaction is the formation of a CuII-O-CuII species, which is more special than it appears at first sight: a single oxo ligand between two CuII centers experiences considerable electron density, and this makes the corresponding complexes reactive and difficult to access. Hence, only a small number of these compounds have been synthesized and characterized unequivocally to date, and as biological relevance was not apparent, they remained unappreciated. However, recently they moved into the spotlight, when CuII-O-CuII cores were proposed as the active species in the challenging oxidation of methane to methanol at the surface of a Cu-grafted zeolite and in the active center of the copper enzyme particulate methane monooxygenase. This Minireview provides an overview of these systems with a special focus on their reactivity and spectroscopic features.
Co-reporter:Dipl.-Chem. Patrick Holze;Dipl.-Chem. Bettina Horn;Dr. Christian Limberg;Dipl.-Chem. Corinna Matlachowski ;Dr. Stefan Mebs
Angewandte Chemie International Edition 2014 Volume 53( Issue 10) pp:2750-2753
Publication Date(Web):
DOI:10.1002/anie.201308270
Abstract
The greenhouse gas sulfur hexafluoride is the common standard example in the literature of a very inert inorganic small molecule that is even stable against O2 in an electric discharge. However, a reduced β-diketiminate nickel species proved to be capable of converting SF6 into sulfide and fluoride compounds at ambient standard conditions. The fluoride product complex features an unprecedented [NiF]+ unit, where the Ni atom is only three-coordinate, while the sulfide product exhibits a rare almost linear [Ni(μ-S)Ni]2+ moiety. The reaction was monitored applying 1H NMR, IR and EPR spectroscopic techniques resulting in the identification of an intermediate nickel complex that gave insight into the mechanism of the eight-electron reduction of SF6.
Co-reporter:Dr. Peter Haack;Dr. Christian Limberg
Angewandte Chemie 2014 Volume 126( Issue 17) pp:4368-4380
Publication Date(Web):
DOI:10.1002/ange.201309505
Abstract
Die Forschung zur O2-Aktivierung an ligierten CuI-Ionen wird durch ihre biologische Relevanz und die Suche nach effizienten Oxidationskatalysatoren getrieben. Ein in diesem Zusammenhang selten beobachteter Reaktionspfad ist die ungewöhnliche Bildung von CuII-O-CuII-Einheiten: Ein einzelner, zwischen zwei CuII-Ionen gebundener Oxoligand erfährt eine beträchtliche Elektronendichte, die Cu2O-Komplexe sehr reaktiv und somit nur schwer zugänglich macht. Daher – und auch wegen der scheinbar fehlenden biologischen Bedeutung – wurden bisher nur wenige solche Verbindungen synthetisiert und charakterisiert. Kürzlich jedoch wurden CuII-O-CuII-Einheiten als aktive Spezies für die selektive Oxidation von Methan zu Methanol sowohl auf der Oberfläche eines kupferhaltigen Zeoliths als auch im aktiven Zentrum der Methanmomooxygenase diskutiert. Dieser Kurzaufsatz soll einen Überblick über den aktuellen Wissensstand bezüglich niedermolekularer Systeme bieten, mit einem Fokus auf ihren spektroskopischen Eigenschaften und Reaktivitäten.
Co-reporter:Dipl.-Chem. Patrick Holze;Dipl.-Chem. Bettina Horn;Dr. Christian Limberg;Dipl.-Chem. Corinna Matlachowski ;Dr. Stefan Mebs
Angewandte Chemie 2014 Volume 126( Issue 10) pp:2788-2791
Publication Date(Web):
DOI:10.1002/ange.201308270
Abstract
Das Klimagas Schwefelhexafluorid ist in der Literatur ein viel verwendetes Standardbeispiel für ein sehr reaktionsträges anorganisches Molekül, das selbst durch Disauerstoff in einer elektrischen Entladung nicht angegriffen wird. Dennoch sind reduzierte β-Diketiminato-Nickel-Komplexe unter Standardbedingungen in der Lage, SF6 in Fluorido- und Sulfido-Verbindungen zu überführen. Das Fluorido-Produkt besitzt dabei eine präzedenzlose [NiF]+-Einheit, bei der das Nickelatom nur dreifach koordiniert ist. Der Sulfido-Komplex beinhaltet eine seltene nahezu lineare [Ni(μ-S)Ni]2+-Anordnung. Die Reaktion wurde mittels 1H-NMR-, IR- und ESR-spektroskopischen Methoden untersucht. Diese führten zur Identifizierung einer Zwischenstufe und gewährten so einen Einblick in den Mechanismus der 8-Elektronen-Reduktion von SF6 .
Co-reporter:Ramona Metzinger;Dr. Serhiy Demeshko;Dr. Christian Limberg
Chemistry - A European Journal 2014 Volume 20( Issue 16) pp:4721-4735
Publication Date(Web):
DOI:10.1002/chem.201304535
Abstract
A novel redox-active ligand, H4Ph2SLAP (1) which was designed to be potentially pentadentate with an O,N,S,N,O donor set is described. Treatment of 1 with two equivalents of potassium hydride gave access to octametallic precursor complex [H2Ph2SLAPK2(thf)]4 (2), which reacted with FeCl3 to yield iron(III) complex [H2Ph2SLAPFeCl] (3). Employing Fe[N(SiMe3)2]3 for a direct reaction with 1 led to ligand rearrangement through CS bond cleavage and thiolate formation, finally yielding [HLAPFe] (5). Upon exposure to O2, 3 and 5 are oxidized through formal hydrogen-atom abstraction from the ligand NH units to form [Ph2SLSQFeCl] (4) and [LSQFe] (6) featuring two or one coordinated iminosemiquinone moieties, respectively. Mössbauer measurements demonstrated that the iron centers remain in their +III oxidation states. Compounds 3 and 5 were tested with respect to their potential as models for the catechol dioxygenase. Thus, they were treated with 3,5-di-tert-butyl-catechol, triethylamine and O2. It turned out that the iron–catecholate complexes react with O2 in dichloromethane at ambient conditions through CC bond cleavage mainly forming extradiol cleavage products. Intradiol products are only side products and quinone formation becomes negligible. This observation has been rationalized by a dissociation of two donor functions upon coordination of the catecholate.
Co-reporter:Dipl.-Chem. Fabian Schax;Dr. Eckhard Bill;Dr. Christian Herwig;Dr. Christian Limberg
Angewandte Chemie 2014 Volume 126( Issue 47) pp:12955-12959
Publication Date(Web):
DOI:10.1002/ange.201406313
Abstract
The reaction of a tripodal trisilanol with n-butyllithium and CrCl2 results in a dinuclear CrII complex (1), which is capable of cleaving O2 to yield in a unique complex (2) with an asymmetric diamond core composed of two CrIVO units. Magnetic susceptibility data reveal significant exchange coupling of CrII (S=2) in 1 and large zero-field splitting for CrIV (S=1) in 2 owing to strong spin–orbit coupling of the ground state. The CrIVO compound can also be generated using PhIO, and evidence was gathered that although it is the stable product isolated after excessive O2 treatment, it further activates O2 to yield an intermediate species that oxidizes THF or Me-THF. By extensive 18O labeling studies we were able to show, that in the course of this process 18O2 exchanges its label with siloxide O atoms of the ligand via terminal oxido ligands.
Co-reporter:Dipl.-Chem. Fabian Schax;Dr. Eckhard Bill;Dr. Christian Herwig;Dr. Christian Limberg
Angewandte Chemie International Edition 2014 Volume 53( Issue 47) pp:12741-12745
Publication Date(Web):
DOI:10.1002/anie.201406313
Abstract
The reaction of a tripodal trisilanol with n-butyllithium and CrCl2 results in a dinuclear CrII complex (1), which is capable of cleaving O2 to yield in a unique complex (2) with an asymmetric diamond core composed of two CrIVO units. Magnetic susceptibility data reveal significant exchange coupling of CrII (S=2) in 1 and large zero-field splitting for CrIV (S=1) in 2 owing to strong spin–orbit coupling of the ground state. The CrIVO compound can also be generated using PhIO, and evidence was gathered that although it is the stable product isolated after excessive O2 treatment, it further activates O2 to yield an intermediate species that oxidizes THF or Me-THF. By extensive 18O labeling studies we were able to show, that in the course of this process 18O2 exchanges its label with siloxide O atoms of the ligand via terminal oxido ligands.
Co-reporter:Peter Haack ; Anne Kärgel ; Claudio Greco ; Jadranka Dokic ; Beatrice Braun ; Florian F. Pfaff ; Stefan Mebs ; Kallol Ray
Journal of the American Chemical Society 2013 Volume 135(Issue 43) pp:16148-16160
Publication Date(Web):October 17, 2013
DOI:10.1021/ja406721a
We report a complex with a rare CuII–O–CuII structural motif that is stable at room temperature, which allows its in-depth characterization by a variety of spectroscopic methods. Interest in such compounds is fueled by the recent discovery that a CuII–O–CuII species on the surface of Cu-ZSM-5 is capable of oxidizing methane to methanol, and this in turn ties into mechanistic discussions on the methane oxidation at the dicopper site within the particulate methane monooxygenase. For the synthesis of our Cu2O complex we have developed a novel, neutral ligand system, FurNeu, exhibiting two N-(N′,N′-dimethylaminoethyl)(2-pyridylmethyl)amino binding pockets connected by a dibenzofuran spacer. The reaction of FurNeu with CuCl yielded [FurNeu](Cu2(μ-Cl))(CuCl2), 1, demonstrating the geometric potential of the ligand to stabilize Cu–X–Cu moieties. A CuI precursor with weakly coordinating anions was chosen in the next step, namely [Cu(NCCH3)4]OTf, which led to the formation of [FurNeu](Cu(NCCH3))2(OTf)2, 3. Treatment of 3 with O2 or PhIO led to identical green solutions, whose UV–vis spectra were markedly different from the one displayed by [FurNeu](Cu)2(OTf)4, 4, prepared independently from FurNeu and Cu(OTf)2. Further investigations including PhIO consumption experiments, NMR and UV–vis spectroscopy, HR-ESI mass spectrometry, and protonation studies led to the identification of the green product as [FurNeu](Cu2(μ-O))(OTf)2, 5. DOSY NMR spectroscopy confirmed its monomeric character. Over longer periods of time 5 decomposes to give [Cu(picoloyl)2], formed through an oxidative N-dealkylation reaction followed by further oxidation of the ligand. Due to its slow decomposition reaction, all attempts to crystallize 5 failed. However, its structure in solution could be determined by EXAFS analysis in combination with DFT calculations, which revealed a Cu–O–Cu angle that amounts to 105.17°. Moreover, TDDFT calculations helped to rationalize the UV–vis absorptions of 5. The reactivity of complex 5 with 2,4-di-tert-butylphenol, DTBP, was also investigated; the initially formed biphenol product, TBBP, was found to further react in the presence of excessive O2 to yield 2,4,7,9-tetra-tert-butyloxepino[2,3-b]benzofuran, TBOBF, via an intermediate diphenoquinone. It turned out that 5, or its precursor 3, can even be employed as a catalyst for the oxidation of DTBP to TBBP or for the oxidation of TBBP to TBOBF.
Co-reporter:Bettina Horn, Christian Limberg, Christian Herwig and Beatrice Braun
Chemical Communications 2013 vol. 49(Issue 93) pp:10923-10925
Publication Date(Web):17 Oct 2013
DOI:10.1039/C3CC45407J
The β-diketiminato nickel(I) complex K2[LtBuNiI(N22−)NiILtBu] reacts with CO2via reductive disproportionation to form CO and CO32− containing products, whereas after employment of the NiI precursor [LtBuNiI(N2)NiILtBu] reductive coupling of CO2 was observed giving an oxalate bridged dinickel(II) complex. The addition of KC8 to the carbonate and oxalate compounds formed leads to the regeneration of the initial NiI complexes in an N2 atmosphere, thus closing synthetic cycles.
Co-reporter:Dipl.-Chem. Denise Pinkert;Dr. Serhiy Demeshko;Dipl.-Chem. Fabian Schax;Dr. Beatrice Braun;Dr. Franc Meyer;Dr. Christian Limberg
Angewandte Chemie International Edition 2013 Volume 52( Issue 19) pp:5155-5158
Publication Date(Web):
DOI:10.1002/anie.201209650
Co-reporter:Subrata Kundu;Florian Felix Pfaff;Enrico Miceli;Dr. Ivelina Zaharieva;Dr. Christian Herwig;Dr. Shenglai Yao;Dr. Erik R. Farquhar;Dr. Uwe Kuhlmann;Dr. Eckhard Bill;Dr. Peter Hildebrt;Dr. Holger Dau;Dr. Matthias Driess;Dr. Christian Limberg;Dr. Kallol Ray
Angewandte Chemie International Edition 2013 Volume 52( Issue 21) pp:5622-5626
Publication Date(Web):
DOI:10.1002/anie.201300861
Co-reporter:Dipl.-Chem. Denise Pinkert;Dr. Serhiy Demeshko;Dipl.-Chem. Fabian Schax;Dr. Beatrice Braun;Dr. Franc Meyer;Dr. Christian Limberg
Angewandte Chemie 2013 Volume 125( Issue 19) pp:5260-5263
Publication Date(Web):
DOI:10.1002/ange.201209650
Co-reporter:Henrike Gehring;Ramona Metzinger;Christian Herwig;Julia Intemann;Dr. Sjoerd Harder;Dr. Christian Limberg
Chemistry - A European Journal 2013 Volume 19( Issue 5) pp:1629-1636
Publication Date(Web):
DOI:10.1002/chem.201203201
Abstract
After the lithiation of PYR-H2 (PYR2−=[{NC(Me)C(H)C(Me)NC6H3(iPr)2}2(C5H3N)]2−), which is the precursor of an expanded β-diketiminato ligand system with two binding pockets, its reaction with [NiBr2(dme)] led to a dinuclear nickel(II)–bromide complex, [(PYR)Ni(μ-Br)NiBr] (1). The bridging bromide ligand could be selectively exchanged for a thiolate ligand to yield [(PYR)Ni(μ-SEt)NiBr] (3). In an attempt to introduce hydride ligands, both compounds were treated with KHBEt3. This treatment afforded [(PYR)Ni(μ-H)Ni] (2), which is a mixed valent NiIμ-HNiII complex, and [(PYR-H)Ni(μ-SEt)Ni] (4), in which two tricoordinated NiI moieties are strongly antiferromagnetically coupled. Compound 4 is the product of an initial salt metathesis, followed by an intramolecular redox process that separates the original hydride ligand into two electrons, which reduce the metal centres, and a proton, which is trapped by one of the binding pockets, thereby converting it into an olefin ligand on one of the NiI centres. The addition of a mild acid to complex 4 leads to the elimination of H2 and the formation of a NiIINiII compound, [(PYR)Ni(μ-SEt)NiOTf] (5), so that the original NiII(μ-SEt)NiIIX core of compound 3 is restored. All of these compounds were fully characterized, including by X-ray diffraction, and their molecular structures, as well as their formation processes, are discussed.
Co-reporter:Bettina Horn, Christian Limberg, Christian Herwig, Michael Feist and Stefan Mebs
Chemical Communications 2012 vol. 48(Issue 66) pp:8243-8245
Publication Date(Web):12 Jul 2012
DOI:10.1039/C2CC33846G
Reaction of a nickel(0) carbonyl complex, K2[LtBuNiCO]2, with N2O generates a cyclic carbonate compound composed of six [NiII(CO3)K]+ units. The same product can also be obtained using O2 as the oxidant in a solid-state/gas reaction. These conversions represent unique examples of a nickel-bound CO oxidation by N2O and O2, respectively.
Co-reporter:Claudia Köthe, Ramona Metzinger, Christian Herwig, and Christian Limberg
Inorganic Chemistry 2012 Volume 51(Issue 18) pp:9740-9747
Publication Date(Web):September 5, 2012
DOI:10.1021/ic301066x
The reaction of [LtBuNi(OEt2)] (where LtBu = [HC(CtBuNC6H3(iPr)2)2]−) with phenylhydrazine leads to the phenylhydrazido(1−) complex [LtBuNi(η2-NPhNH2)] (1) with concomitant formation of H2. Treatment of 1 with potassium graphite in the presence of crown ether again leads to H2 evolution and affords the heterobimetallic complex [LtBuNi(μ-η2:η2-NPhNH)]K(18-crown-6) containing the doubly deprotonated phenylhydrazido(2−) ligand. 1 can be converted into a phenyldiazenido complex [LtBuNi(η1-NNPh)] in the course of a dehydrogenation reaction employing 1,2-diisopropylazo dicarboxylate (DIAD) as the oxidant.
Co-reporter:Aline Arnold, Christian Limberg, and Ramona Metzinger
Inorganic Chemistry 2012 Volume 51(Issue 22) pp:12210-12217
Publication Date(Web):October 30, 2012
DOI:10.1021/ic301391s
A novel chiral ligand system L containing one pyridyl and two imidazolyl donor functions has been synthesized and investigated with respect to its CuI and CuII coordination chemistry. Reaction with [Cu(MeCN)4]PF6 and [Cu(MeCN)4]OTf led to the dimeric complexes [LCu]2X2 (1, X = PF6; 2, X = OTf) with the ligands L in different configurations (R,S). The ligand matrix formed in these complexes can also host a CuII ion instead of two CuI ions so that mixed crystals of [L2Cu]X2 and [LCu]2X2 can be produced. The pure compounds [L2Cu]X2 (3, X = PF6; 4, X = OTf) can be obtained by treatment of 1 and 2 with O2 in acetonitrile, respectively. From the corresponding solution 3 crystallizes with the two L molecules in different configurations, while 4 crystallizes with the ligands in (S,S) or (R,R) configurations, respectively. Crystals containing the analogous diastereomers of 3 were obtained, besides those isolated previously, when this compound was synthesized by reaction of 1 with AgPF6. On treating 2 with O2 as the oxidant in acetonitrile, besides formation of 4, additional evidence for oxygenation of L to Lox, where one of the original phenyl units corresponds to a phenolate function, was found: The dinuclear complex [LoxCu(OH)(OTf)CuL](OTf) (5) was isolated as the final product of O2 activation and conversion, which resembles the one of tyrosinase. In acetonitrile 5 reacts further to give 4 and [Lox2Cu2](OTf)2 (6), and hence, product mixtures are obtained. In CH2Cl2 decomposition can be avoided, and hence, changing the solvent from acetonitrile to CH2Cl2 leads to selective formation of 5.
Co-reporter:Fabian Schax, ;Clemens Mügge
European Journal of Inorganic Chemistry 2012 Volume 2012( Issue 29) pp:4661-4668
Publication Date(Web):
DOI:10.1002/ejic.201200294
Abstract
Reaction of the siloxane-diols HO(Ph2SiO)2H, 1,1,3,3-tetraphenyldisiloxane-1,3-diol (L1H2), HO(iPr2SiO)2H, 1,1,3,3-tetraisopropyldisiloxane-1,3-diol (L2H2), and HO(Ph2SiO)3H, 1,1,3,3,5,5-hexaphenyltrisiloxane-1,5-diol (L3H2) with two equivalents of CunMesn led to octanuclear compounds [Cu8L14] (1), [Cu8L24] (2), and [Cu8L12L*2] [L* = O(Ph2SiO)4] (3), which were characterized by single-crystal X-ray diffraction analysis as well as by solution NMR spectroscopy. The crystal structures revealed that all the compounds are composed of Cu4O4 moieties featuring short Cu–Cu distances that can be discussed in terms of cuprophilic interactions. Two such units are linked via four disiloxane units in 1 and 2, in 3 they are connected by two equivalents of (L1)2–. Formation of 3 required disproportion of a trisiloxane-1,5-diolate to give a disiloxane-1,3-diolate and a tetrasiloxane-1,7-diolate, and thus indicates that siloxane units are not inert under the synthetic conditions employed. NMR spectroscopic investigations of the structure of 1 in solution revealed an equilibrium, presumably with [Cu4L22] fragments. Crystallization of a structural isomer of 3 indicates a shallow potential energy surface for this compound.
Co-reporter:Dipl.-Chem. Carolin Tschersich;Dr. Christian Limberg;Dr. Stefan Roggan;Dr. Christian Herwig;Dr. Nikolaus Ernsting;Dr. Sergey Kovalenko ;Dr. Stefan Mebs
Angewandte Chemie International Edition 2012 Volume 51( Issue 20) pp:4989-4992
Publication Date(Web):
DOI:10.1002/anie.201200848
Co-reporter:Dipl.-Chem. Madleen Sallmann;Dr. Inke Siewert;Dipl.-Chem. Lea Fohlmeister;Dr. Christian Limberg;Dr. Christina Knispel
Angewandte Chemie International Edition 2012 Volume 51( Issue 9) pp:2234-2237
Publication Date(Web):
DOI:10.1002/anie.201107345
Co-reporter:Dipl.-Chem. Carolin Tschersich;Dr. Christian Limberg;Dr. Stefan Roggan;Dr. Christian Herwig;Dr. Nikolaus Ernsting;Dr. Sergey Kovalenko ;Dr. Stefan Mebs
Angewandte Chemie 2012 Volume 124( Issue 20) pp:5073-5077
Publication Date(Web):
DOI:10.1002/ange.201200848
Co-reporter:Dipl.-Chem. Madleen Sallmann;Dr. Inke Siewert;Dipl.-Chem. Lea Fohlmeister;Dr. Christian Limberg;Dr. Christina Knispel
Angewandte Chemie 2012 Volume 124( Issue 9) pp:2277-2280
Publication Date(Web):
DOI:10.1002/ange.201107345
Co-reporter:Dr. Ann-Katrin Jungton;Dr. Christian Herwig;Dr. Thomas Braun;Dr. Christian Limberg
Chemistry - A European Journal 2012 Volume 18( Issue 32) pp:10009-10013
Publication Date(Web):
DOI:10.1002/chem.201201295
Abstract
1H NMR exchange spectroscopy of a reaction mixture of [Cp*Ir(H)4] (1; Cp*=1,2,3,4,5-pentamethylcyclopentadienyl) and ammonia suggests an exchange of hydrogen atoms between the hydrido ligands and ammonia. Treatment of 1 with ND3 led to an H/D exchange between ND3 and the hydrido ligands of 1. Subsequent studies showed that photolysis of 1 isolated in frozen argon matrices leads to the formation of the iridium compounds [Cp*Ir(H)2] (2) and [Cp*Ir(H)3] (4), as it was confirmed by IR spectroscopy. In the presence of water the aqua complex [Cp*Ir(H)2(OH2)] (3) was generated simultaneously. Accordingly, photolysis of 1 in an argon matrix doped with ammonia gave rise to the ammine complex [Cp*Ir(H)2(NH3)] (5). IR assignments were supported by calculations of the gas-phase IR spectra of 1–5 by DFT methods.
Co-reporter:Peter Haack, Christian Limberg, Thomas Tietz and Ramona Metzinger
Chemical Communications 2011 vol. 47(Issue 22) pp:6374-6376
Publication Date(Web):06 May 2011
DOI:10.1039/C1CC11518A
The first structural characterisation of a copper–carbondisulfide complex revealed a hitherto unknown binding mode for CS2: it interacts with two metal centres (CuI) simultaneously via both CS π bonds. DFT calculations showed that complex formation occurs mainly due to a donation of electron density from the copper centres into the CS π* orbitals.
Co-reporter:Christina Knispel, Christian Limberg and Carolin Tschersich
Chemical Communications 2011 vol. 47(Issue 38) pp:10794-10796
Publication Date(Web):05 Sep 2011
DOI:10.1039/C1CC14616E
Reaction of [Cp2MoH2] with bismuth allyloxide, [Bi{OCH(CH3)CHCH2}3], gave rise to an extended octanuclear complex wherein two cyclic Mo2Bi2 units composed of four Mo–Bi bonds are linked by a Bi–Bi bond. The fact that the construction of such an assembly could be accomplished only in the case of a monomethylation of the parent allyl residue demonstrates a subtle substituent effect.
Co-reporter:Peter Haack ; Christian Limberg ; Kallol Ray ; Beatrice Braun ; Uwe Kuhlmann ; Peter Hildebrandt ;Christian Herwig
Inorganic Chemistry 2011 Volume 50(Issue 6) pp:2133-2142
Publication Date(Web):February 22, 2011
DOI:10.1021/ic101249k
Investigations concerning the system β-diketiminato-CuI/O2 have revealed valuable insights that may be discussed in terms of the behavior of mononuclear oxygenases containing copper. On the other hand nature also employs dinuclear Cu enzymes for the activation of O2. With this background the ligand system [Me2C6H3Xanthdim]2− containing two parallel β-diiminato binding sites linked by a xanthene backbone with 2,3-dimethylphenyl residues at the diiminato units was investigated with respect to its copper coordination chemistry. The diimine [Me2C6H3Xanthdim]H2 was treated with CuOtBu in the presence of acetonitrile, PPh3, and PMe3 to yield the corresponding complexes [Me2C6H3Xanthdim](Cu(L))2 (L = CH3CN, 1, PPh3, 2, and PMe3, 3) that proved to be stable and were fully characterized. Single crystal X-ray diffraction analyses performed for the three complexes showed that considerable steric crowding within the binding pockets of 2 leads to a very long Cu−Cu distance while the structures of 1 and 3 are relaxed. Compounds 2 and 3 are relatively robust toward air, whereas 1 is very sensitive and quantitatively reacts with O2 at room temperature (r.t.) within less than 2 min to give intractable compounds. At low temperatures the formation of a green intermediate was observed that was identified as a CuII−O−CuII species spectroscopically and chemically. This finding is of relevance also in the context of the results obtained testing 1 as a catalyst for phenol oxidation using O2: 1 efficiently catalyzes phenol coupling, while there was no evidence for any oxygenation reactions occurring.
Co-reporter:Dipl.-Chem. Thomas Tietz;Dr. Christian Limberg;Dr. Reinhard Stößer ;Dr. Burkhard Ziemer
Chemistry - A European Journal 2011 Volume 17( Issue 36) pp:10010-10020
Publication Date(Web):
DOI:10.1002/chem.201100343
Abstract
A series of complexes of the type [(TpR1,R2)M(X)] (Tp=trispyrazolylborato) with R1/R2 combinations Me/tBu, Ph/Me, iPr/iPr, Me/Me and for M=Mn or Fe coordinating [PzMe,tBu]− (Pz=pyrazolato) or Cl− as co-ligand X has been synthesised. Although the chloride complexes were very unreactive and stable in air, the pyrazolato series was far more reactive in contact with oxidants like O2 and tBuOOH. The [(TpR1,R2)M(PzMe,tBu)] complexes proved to be active pre-catalysts for the oxidation of cyclohexene with tBuOOH, reaching turnover frequencies (TOFs) ranging between moderate and good in comparison to other manganese catalysts. Cyclohexene-3-one and cyclohexene-3-ol were always found to represent the main products, with cyclohexene oxide occasionally formed as a side product. The ratios of the different oxidation products varied with the reaction conditions: in the case of a peroxide/alkene ratio of 4:1, considerably more ketone than alcohol was obtained and cyclohexene oxide formation was almost negligible, whereas a ratio of 1:10 led to a significant increase of the alcohol proportion and to the formation of at least small amounts of the epoxide. Pre-treatment of the dissolved [(TpR1,R2)M(PzMe,tBu)] pre-catalysts with O2 led to product distributions and TOFs that were very similar to those found in the absence of O2, so that it may be argued that tBuOOH and O2 both lead to the same active species. The results of EPR spectroscopy and ESI-MS suggest that the initial product of the reaction of [(TpMe,Me)Mn(PzMe,tBu)] with O2 contains a MnIII(O)2MnIV core. Prolonged exposure to O2 leads to a different dinuclear complex containing three O-bridges and resulting in different TOFs/product distributions. Analogous findings were made for other complexes and formation of these overoxidised products may explain the deviation of the catalytic performances if the reactions are carried out in an O2 atmosphere.
Co-reporter:Dipl.-Chem. Bettina Horn;Dr. Christian Limberg;Dr. Christian Herwig ;Dr. Stefan Mebs
Angewandte Chemie 2011 Volume 123( Issue 52) pp:12829-12833
Publication Date(Web):
DOI:10.1002/ange.201105281
Co-reporter:Dipl.-Chem. Bettina Horn;Dr. Christian Limberg;Dr. Christian Herwig ;Dr. Stefan Mebs
Angewandte Chemie International Edition 2011 Volume 50( Issue 52) pp:12621-12625
Publication Date(Web):
DOI:10.1002/anie.201105281
Co-reporter:Christina Knispel and Christian Limberg
Organometallics 2011 Volume 30(Issue 14) pp:3701-3703
Publication Date(Web):June 21, 2011
DOI:10.1021/om2004223
The reaction of [LBiNMe2], I (L = [1,8-C10H6(NSiMe3)2]2–), with [MeCp2Mo═O] results in a coordination of the Mo═O group at the Bi center, which in turn triggers C–H activation at the methylcyclopentadienyl ligand through complex-induced proximity effects. This leads to a heterobimetallic molybdenum/bismuth compound, wherein the two metal centers are linked by an oxide bridge as well as by a σ,π-binding [μ-η5:η1-CH2(C5H4)]2– ligand.
Co-reporter:C. Gunnar Werncke;Dr. Christian Limberg;Christina Knispel;Ramona Metzinger ;Dr. Beatrice Braun
Chemistry - A European Journal 2011 Volume 17( Issue 10) pp:2931-2938
Publication Date(Web):
DOI:10.1002/chem.201002890
Abstract
By employing the 2,2′-thiobis(2,4-di-tert-butylphenolate) ligand (SL2−) a novel oxovanadium(V) complex, (PPh4)2[SLV(O)(μ2-O)2V(O)SL] (1), was synthesised that exhibits haloperoxidase activity: on addition of H2O2 a sequence of successive peroxide formation and intramolecular thioether oxidation events (sulfoxide and sulfone) led to a mixture of five products, which were all identified unambiguously, partly through an independent synthesis and characterisation. It was shown that internal thioether oxidation proceeds through peroxide formation, but the sulfoxidation of external thioether functions requires further activation of the peroxide function by protons or alkyl cations. Consistently, the employment of tBuOOH instead of H2O2 led to a very active system for the catalytic sulfoxidation of thioethers.
Co-reporter:C. Gunnar Werncke;Dr. Christian Limberg;Dr. Christina Knispel ;Dr. Stefan Mebs
Chemistry - A European Journal 2011 Volume 17( Issue 43) pp:12129-12135
Publication Date(Web):
DOI:10.1002/chem.201101442
Abstract
On the basis that thiacalix[4]arene (H4T4A) complex (PPh4)2[H2T4A(VO2)]2 (Ia) was found to be an adequate functional model for surface species occurring on vanadium oxide based catalysts and itself catalyses the oxidative dehydrogenation (ODH) of alcohols, an analogue containing 2,2′-thiobis(2,4-di-tert-butylphenolate), SL2−, as ligand, namely, (PPh4)2[SLVO2]2 (II) was investigated in the same context. Despite the apparent similarity of Ia and II, studies on II revealed several novel insights, which are also valuable in connection with surfaces of vanadia catalysts: 1) For Ia and II similar turnover numbers (TONs) were found for the ODH of activated alcohols, which indicates that the additional OH units inherent to Ia do not contribute particularly to the activity of this complex, for instance, through prebinding of the alcohol. 2) On dissolution II enters into an equilibrium with a monomeric form, which is the predominant species in solution; nevertheless, ODH proceeds exclusively at the dimeric form, and this stresses the need for cooperation of two vanadium centres. 3) By omitting O2 from the system during the oxidation of 9-fluorenol, the reduced form of the catalyst could be isolated and fully characterised (including single-crystal X-ray analysis). The corresponding intermediate had been elusive in case of thiacalixarene system Ia. 4) Reoxidation was found to proceed via a peroxide intermediate that also oxidises one alcohol equivalent. As the peroxide can also perform mono- and dioxygenation of the thioether group in II, after a number of turnovers the active catalyst contains a sulfone group. The reduced form of this ultimate catalyst was also isolated and structurally characterised. Possible implications of 1)–4) for the function of heterogeneous vanadia catalysts are discussed.
Co-reporter:Stefan Pfirrmann ; Christian Limberg ; Christian Herwig ; Christina Knispel ; Beatrice Braun ; Eckhard Bill ;Reinhard Stösser
Journal of the American Chemical Society 2010 Volume 132(Issue 39) pp:13684-13691
Publication Date(Web):September 9, 2010
DOI:10.1021/ja106266v
An investigation concerning the stepwise reduction of the β-diketiminato nickel(II) hydride dimer [LNi(μ-H)2NiL], 1 (L = [HC(CMeNC6H3(iPr)2)2]−), has been carried out. While the reaction with one equivalent of potassium graphite, KC8, led to the mixed valent NiI/NiII complex K[LNi(μ-H)2NiL], 3, treatment of 1 with two equivalents of KC8 surprisingly yielded in the trinuclear complex K2[LNi(μ-H)2Ni(μ-H)2NiL], 4, in good yields. The Ni3H4 core contains one NiII and two NiI centers, which are antiferromagnetically coupled so that a singlet ground state results. 4 represents the first structurally characterized molecular compound with three nickel atoms bridged by hydride ligands, and it shows a very interesting chemical behavior: Single-electron oxidation yields in the NiII2NiI compound K[LNi(μ-H)2Ni(μ-H)2NiL], 5, and treatment with CO leads to the elimination of H2 with formation of the carbonyl complex K2[LNi(CO)]2, 6. Beyond that, it could be shown that 4 undergoes H/D exchange with deuterated solvents and the deuteride-compound 4-D4 reacts with H2 to give back 4. The crystal structures of the novel compounds 3−6 have been determined, and their electronic structures have been investigated by EPR and NMR spectroscopy, magnetic measurements, and DFT calculations.
Co-reporter:Christina Knispel ; Christian Limberg ;Burkhard Ziemer
Inorganic Chemistry 2010 Volume 49(Issue 9) pp:4313-4318
Publication Date(Web):April 12, 2010
DOI:10.1021/ic100171f
Bismuth allyloxides, [Bi(OR)3] with R = CH2CH═CH2, CH(CH3)CH═CH2, C(CH3)2CH═CH2, and CH2CH═C(CH3)2, can be prepared by alcoholysis of [Bi(OtBu)3] and, in some cases, also via salt metathesis reactions starting from BiCl3 and sodium allylates. They are readily soluble in common organic solvents, and NMR spectroscopic investigations do not provide any hint to aggregated species or any equilibria in solution. The majority of the compounds also proved volatile enough to be purified by sublimation. Crystal structure analyses, however, provided evidence for a high degree of aggregation in the solid state, which leads to large rings and chains as structural motifs.
Co-reporter:Marc Ostermeier Dipl.-Chem.;Marie-Anne Berlin Dipl.-Chem.;RobertM. Meudtner Dr.;Serhiy Demeshko Dr.;Franc Meyer Dr. Dr.;Stefan Hecht Dr.
Chemistry - A European Journal 2010 Volume 16( Issue 33) pp:10202-10213
Publication Date(Web):
DOI:10.1002/chem.201000721
Abstract
2,6-Bis(1,2,3-triazol-4-yl)pyridine (btp) ligands with substitution patterns ranging from strongly electron-donating to strongly electron-accepting groups, readily prepared by means of Cu-catalyzed 1,3-dipolar cycloaddition (the “click” reaction), were investigated with regard to their complexation behavior, and the properties of the resulting transition-metal compounds were compared. Metal–btp complexes of 1:1 stoichiometry, that is, [Ru(btp)Cl2(dmso)] and [Zn(btp)Br2], could be isolated and were crystallographically characterized: they display octahedral and trigonal-bipyramidal coordination geometries, respectively, and exhibit high aggregation tendencies due to efficient π–π stacking leading to low solubilities. Metal–btp complexes of 1:2 stoichiometry, that is, [Fe(btp)2]2+ and [Ru(btp)2]2+, could also be synthesized and their metal centers show the expected octahedral coordination spheres. The iron compounds exhibit quite a complex magnetic behavior in the solid state including spin crossover near room temperature, and hysteresis and locking into high-spin states on tempering at 400 K, depending on the substituents on the btp ligands. Cyclic voltammetry studies of [Ru(btp)2]2+ reveal strong modulation of the oxidation potentials by more than 0.6 V and a clear linear correlation to the Hammett constant (σpara) of the substituent at the pyridine core. Isothermal titration calorimetry was used to measure the thermodynamics of the FeII–btp complexation process and enabled accurate determination of the complexation enthalpies, which display a linear relationship with the σpara values for the terminal phenyl substituents. Detailed NMR spectroscopic studies finally revealed that in the case of FeII complexation, dynamics are rapid for all investigated btp derivatives in acetonitrile, while replacing FeII by RuII or changing the solvent to dichloromethane effectively slows down ligand exchange. The results nicely demonstrate the utility of substituent parameters, originally developed for linear free-energy relationships to explain reactivity in organic reactions, in coordination chemistry, and to illustrate the potential to custom-design btp ligands and complexes thereof with predictable properties. The fast equilibration of the [Fe(btp)2]2+ complexes together with their tunable stability and interesting magnetic properties should enable the design of dynamic metallosupramolecular materials with advantageous properties.
Co-reporter:Christian Ohde Dr.
Chemistry - A European Journal 2010 Volume 16( Issue 23) pp:6892-6899
Publication Date(Web):
DOI:10.1002/chem.201000171
Abstract
Silsesquioxane dioxovanadate(V) complexes were investigated with respect to their potential as a catalyst for the oxidative dehydrogenation of alcohols with O2 as an oxidant. The turnover frequencies determined were comparatively low, but during the oxidation of cinnamic alcohol an increase in activity was observed in the course of the process, which was inspected more closely. It turned out that during the oxidation of cinnamic alcohol, not only was the aldehyde formed but also cinnamic acid, which in turn reacts with the silsesquioxane complex employed to give NBu4[O2V(O2CC2H2Ph)2], which can also be obtained from NBu4VO3 and cinnamic acid and represents a far more active catalyst, not only for cinnamic alcohol but also for other activated alcohols and hydrocarbons. The rate-determining step of the conversion corresponds to an hydrogen-atom abstraction from the CH units, as shown by the determination of the kinetic isotope effect in case of 9-hydroxyfluorene, and the reoxidation of the reduced catalyst proceeds via a peroxo intermediate, which is also capable of oxidizing one alcohol equivalent. Furthermore the influence of the organic residues at the carboxylate ligands on the catalyst performance was investigated, which showed that the activity increases with decreasing pKs value. Moreover, it was found that during the oxidation the catalyst slowly decomposes, but can be regenerated by addition of excessive carboxylic acid.
Co-reporter:Dr. Shenglai Yao;Dr. Christian Herwig;Dr. Yun Xiong;Dr. Anna Company;Dr. Eckhard Bill;Dr. Christian Limberg;Dr. Matthias Driess
Angewandte Chemie 2010 Volume 122( Issue 39) pp:7208-7212
Publication Date(Web):
DOI:10.1002/ange.201001914
Co-reporter:Rafael Schiwon, Christina Knispel and Christian Limberg
Organometallics 2010 Volume 29(Issue 7) pp:1670-1674
Publication Date(Web):March 1, 2010
DOI:10.1021/om100011r
The reaction of rhenocene hydride with bismuth alkoxides leads to complexes containing Re−Bi metal bonds with concomitant formation of the corresponding alcohols. Hence, compounds of the type [Cp2Re−BiR2] can be obtained from the monoalkoxide [(MeO)Bi(o-tol)2]n and also from the trialkoxide [Bi{OCH(CF3)2}3(thf)]2, for which in principle also multiple substitution reactions would have been possible, but are prohibited by the electron-withdrawing character of the hexafluoroisopropyl groups. Correspondingly, the trialkoxide [Bi(OtBu)3] containing electron-rich tert-butyl groups does lead to multiple substitution events: First of all it reacts with two equivalents of [Cp2ReH] to give the intermediate [(Cp2Re)2Bi(OtBu)], which could not be isolated, since it undergoes an intramolecular alcohol elimination via CpC−H bond cleavage. This results in the complex [CpRe(μ-η5,η1-C5H4)Bi−ReCp2], featuring a bent Bi−C bond so that one deprotonated Cp ligand bridges a Bi−Re metal bond. All compounds have been fully characterized, and their crystal structures are discussed.
Co-reporter:Christian Limberg ;Matthias Driess
ChemCatChem 2010 Volume 2( Issue 7) pp:711-712
Publication Date(Web):
DOI:10.1002/cctc.201000200
No abstract is available for this article.
Co-reporter:Holger Dau Dr. Dr.;Tobias Reier;Marcel Risch;Stefan Roggan Dr.;Peter Strasser Dr.
ChemCatChem 2010 Volume 2( Issue 7) pp:724-761
Publication Date(Web):
DOI:10.1002/cctc.201000126
Abstract
Striving for new solar fuels, the water oxidation reaction currently is considered to be a bottleneck, hampering progress in the development of applicable technologies for the conversion of light into storable fuels. This review compares and unifies viewpoints on water oxidation from various fields of catalysis research. The first part deals with the thermodynamic efficiency and mechanisms of electrochemical water splitting by metal oxides on electrode surfaces, explaining the recent concept of the potential-determining step. Subsequently, novel cobalt oxide-based catalysts for heterogeneous (electro)catalysis are discussed. These may share structural and functional properties with surface oxides, multinuclear molecular catalysts and the catalytic manganese–calcium complex of photosynthetic water oxidation. Recent developments in homogeneous water-oxidation catalysis are outlined with a focus on the discovery of mononuclear ruthenium (and non-ruthenium) complexes that efficiently mediate O2 evolution from water. Water oxidation in photosynthesis is the subject of a concise presentation of structure and function of the natural paragon—the manganese–calcium complex in photosystem II—for which ideas concerning redox-potential leveling, proton removal, and OO bond formation mechanisms are discussed. The last part highlights common themes and unifying concepts.
Co-reporter:Dr. Shenglai Yao;Dr. Christian Herwig;Dr. Yun Xiong;Dr. Anna Company;Dr. Eckhard Bill;Dr. Christian Limberg;Dr. Matthias Driess
Angewandte Chemie International Edition 2010 Volume 49( Issue 39) pp:7054-7058
Publication Date(Web):
DOI:10.1002/anie.201001914
Co-reporter:Dirk F.-J. Piesik ; Peter Haack ; Sjoerd Harder
Inorganic Chemistry 2009 Volume 48(Issue 23) pp:11259-11264
Publication Date(Web):October 28, 2009
DOI:10.1021/ic9017536
Reaction of the diprotic ligand [Xanthdim]H2 (a ligand system where two adjacent β-dialdimine units are linked by a xanthyl backbone) with 2 equiv of potassium hydride or benzylcesium gave access to bimetallic alkali metal complexes. These complexes were structurally characterized by X-ray diffraction, which showed that the β-diiminato units are orientated in a W-conformation. Treatment with 2 equiv of Mg[N(SiMe3)2]2(THF)2 led to the formation of the heteroleptic complex [Xanthdim][MgN(SiMe3)2(THF)]2, that crystallized as a highly strained monomer. The heteroleptic Mg complex is remarkably stable against ligand exchange but is not active in CO2/cyclohexene oxide copolymerization. Reaction with Ca[N(SiMe3)2]2(THF)2 gave the homoleptic complex [Xanthdim][Ca(THF)]. Both alkaline-earth metal complexes display considerable distortions in their solid state structure.
Co-reporter:Marit Wagner, Christian Limberg
Inorganica Chimica Acta 2009 Volume 362(Issue 13) pp:4809-4812
Publication Date(Web):15 October 2009
DOI:10.1016/j.ica.2009.07.007
The coordination chemistry of a chiral tripodal ligand L containing pyridyl, imidazolyl and pyrazolyl donor functions has been investigated in combination with zinc(II). While the reaction of a racemic mixture of L with ZnX2 (X = Cl, Br) leads to complexes LZnX2 with tetrahedrally coordinated zinc centres (the pyridyl donor function remains pending), the employment of Zn(ClO4)2 leads to the sandwich complex [L2Zn](ClO4)2 which due to the two possible configurations for L (S and R) occurs in form of two diastereomers (the meso form and the enantiomeric pair SS/RR). The crystal structures of all three compounds are discussed.A tripodal ligand containing pyridyl, imidazolyl and pyrazolyl donor functions has been employed in zinc chemistry. On employment of zinc halides 4-coordinate complexes are formed (see picture) while the replacement of the halides by weaker coordinating anions leads to 6-coordinated sandwich complexes.
Co-reporter:Stefan Pfirrmann Dipl.-Chem., Dr.;Christian Herwig Dr.;Reinhard Stößer Dr. ;Burkhard Ziemer Dr.
Angewandte Chemie International Edition 2009 Volume 48( Issue 18) pp:3357-3361
Publication Date(Web):
DOI:10.1002/anie.200805862
Co-reporter:Stefan Pfirrmann Dipl.-Chem., Dr.;Christian Herwig Dr.;Reinhard Stößer Dr. ;Burkhard Ziemer Dr.
Angewandte Chemie 2009 Volume 121( Issue 18) pp:3407-3411
Publication Date(Web):
DOI:10.1002/ange.200805862
Co-reporter:Stefan Pfirrmann, Shenglai Yao, Burkhard Ziemer, Reinhard Stösser, Matthias Driess and Christian Limberg
Organometallics 2009 Volume 28(Issue 24) pp:6855-6860
Publication Date(Web):October 28, 2009
DOI:10.1021/om9007983
Reaction of [LtBuNiBr] (LtBu = [HC(C(CMe3)NC6H3(iPr)2)2]−) with KC8 in toluene solution yields the complex [LtBuNi(toluene)], 1, where toluene is bound in a η2 mode via a C═C unit of the aromatic ring, as revealed by single-crystal X-ray crystallography and DFT calculations (B3LYP/6-31G*). Performing the same reaction in hexane as the solvent did not lead to a traceable product, so that the β-diketiminato ligand system was changed from LtBu to the less bulky LMe (LMe = [HC(CMeNC6H3(iPr)2)2]−). Reduction of [LMeNiBr]2 with KC8 in diethyl ether led to [LMeNi]2, 2, with intramolecular Ni−aryl interactions, while employment of [LMeNi(μ-Br)2Li(thf)2] as a precursor for a reaction with KC8 in OEt2 led to the complex [LMeNi(μ-Br)Li(thf)2]2, 3. Both complexes 2 and 3 could be fully characterized, also with the aid of XRD, and their reactivity with respect to H2 and N2 was examined. It turned out that they oxidatively add H2 to give the known compound [LMeNi(μ-H)]2, I, while the reaction with N2 provides the dinitrogen complex [(LMeNi)2(N2)], 4.
Co-reporter:Inke Siewert Dr. Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 40) pp:10316-10328
Publication Date(Web):
DOI:10.1002/chem.200901910
Abstract
The active centre of sMMO contains a diiron core ligated by histidine and glutamate residues, which is capable of catalysing a remarkable reaction: the oxidation of methane with O2 yielding methanol. This review describes the results of efforts to prepare low-molecular-weight analogues of this active site directing towards 1) the assignment of the spectroscopic signatures identified for certain intermediates of the sMMO catalytic cycle to structural features and 2) the synthesis of molecular compounds that can mimic the reactivity. The historical development of the model chemistry, which is subdivided into structural and functional mimicking, is illustrated and achievements reached so far are highlighted.
Co-reporter:Marit Wagner Dipl.-Chem., Dr. ;Thomas Tietz Dipl.-Chem.
Chemistry - A European Journal 2009 Volume 15( Issue 22) pp:5567-5576
Publication Date(Web):
DOI:10.1002/chem.200802591
Co-reporter:Christina Knispel, Christian Limberg and Michael Mehring
Organometallics 2009 Volume 28(Issue 2) pp:646-651
Publication Date(Web):December 4, 2008
DOI:10.1021/om8008662
First, the reactivity of molybdocene dihydride toward bismuth siloxides was studied. Depending on the stoichiometry, compounds of the type [Cp2Mo{Bi(OR)2}2] (R = SiMe2tBu, 1, R = SiPh3, 3) or [Cp2Mo(μ-BiOR)2MoCp2] (R = SiMe2tBu, 2t, R = SiPh3, 4t) were obtained. Apart from 4 all of these compounds were characterized via single-crystal X-ray diffraction analyses, which revealed for the compound 2t a trans orientation of the two silanolate groups, similar to that found in case of the corresponding tert-butoxide derivatives. In contrast to those, however, the silanolates 2t and 4t proved to be stable in solution with respect to subsequent intramolecular silanol eliminations, which allowed for detailed NMR spectroscopic investigation. These studies revealed that on dissolution of 2t and 4t, they slowly enter into an equilibrium with isomers containing the two silanolate ligands in a cis configuration, 2c and 4c. However, the cis isomers cannot be isolated from the isomeric mixtures, as the trans isomer always precipitates first. Due to the thermodynamic stability of 2 and 4, these complexes can also be obtained starting from [(η5-Cp)Mo(μ2-η5:η1-C5H4)Bi]2, which contains bent Bi−C bonds and is formed from the tert-butoxide derivative of 2 and 4, [Cp2Mo{Bi(OtBu)}2MoCp2], IIH, by further alcohol elimination as the thermodynamically most favored product in that system: Reacting this compound with the silanols HOSiMe2tBu and HOSiPh3 provides 2t and 4t, respectively, which underlines that these complexes represent the thermodynamic holes within these corresponding systems.
Co-reporter:Christian Limberg
Angewandte Chemie International Edition 2009 Volume 48( Issue 13) pp:2270-2273
Publication Date(Web):
DOI:10.1002/anie.200805977
Co-reporter:Shenglai Yao Dr.;Yun Xiong Dr.;Matthias Vogt Dipl.-Chem.;Hansjörg Grützmacher Dr.;Christian Herwig Dr. Dr.;Matthias Driess Dr.
Angewandte Chemie 2009 Volume 121( Issue 43) pp:8251-8254
Publication Date(Web):
DOI:10.1002/ange.200903772
Co-reporter:Christian Limberg
Angewandte Chemie 2009 Volume 121( Issue 13) pp:2305-2308
Publication Date(Web):
DOI:10.1002/ange.200805977
Co-reporter:Shenglai Yao Dr.;Yun Xiong Dr.;Matthias Vogt Dipl.-Chem.;Hansjörg Grützmacher Dr.;Christian Herwig Dr. Dr.;Matthias Driess Dr.
Angewandte Chemie International Edition 2009 Volume 48( Issue 43) pp:8107-8110
Publication Date(Web):
DOI:10.1002/anie.200903772
Co-reporter:Christian Herwig
Inorganic Chemistry 2008 Volume 47(Issue 8) pp:2937-2939
Publication Date(Web):March 18, 2008
DOI:10.1021/ic8002146
V4O10 can be considered as a molecular model for oxovanadium clusters deposited on the surfaces of corresponding heterogeneous catalysts, and here its generation, matrix isolation, and comprehensive characterization by IR, Raman, and UV/vis spectroscopy are described.
Co-reporter:Stefan Pfirrmann, Christian Limberg and Burkhard Ziemer
Dalton Transactions 2008 (Issue 47) pp:6689-6691
Publication Date(Web):24 Oct 2008
DOI:10.1039/B816136B
The preparation of a novel dinuclear nickel(II) hydride complex and its reactivity that often leads to nickel(I) compounds is described.
Co-reporter:Christian Ohde, Marcus Brandt, Christian Limberg, Jens Döbler, Burckhard Ziemer and Joachim Sauer
Dalton Transactions 2008 (Issue 3) pp:326-331
Publication Date(Web):10 Dec 2007
DOI:10.1039/B714933F
Inspired by surface species proposed to occur on heterogeneous catalysts novel oxovanadium(V) silsesquioxanes were synthesised. Reaction of a T8-silsequioxane containing two geminal OH groups with OV(OiPr)3 led to a dinuclear compound where the geminal disiloxide functions of two silsesquioxanes are bridging two OV(OiPr) moieties (2). Formation of 2 shows that—in contrast to proposals made for silica surfaces—in molecular chemistry a bidentate coordination of geminal siloxides to one vanadium centre is not favourable. With the background that species being doubly anchored to a support have been suggested to play active roles on V2O5/SiO2catalysts an anionic complex has been prepared where a divalent dioxovanadium unit replaces one Si corner of a (RSiO1.5)8 cube (a Si–OH function remains pending) (3). 3 has been intensely investigated by vibrational spectroscopy, and to support assignments not only of the ν(VO) bands but also of the ν(V–O–Si) bands, whose positions are of interest in the area of heterogeneous catalysis, isotopic enrichment studies and DFT calculations have been performed. The corresponding investigations were aided by the synthesis and analysis of a silylated derivative of 3, 4. Moreover, with regard to their potential as structural and spectroscopic models all complexes were characterised by single crystal X-ray diffraction. Finally, 2 and 3 were tested as potential catalysts for the photooxidation of cyclohexane and benzene with O2. While 2 shows a slightly higher activity than vanadylacetylacetonate, 3 leads to significantly increased turnover numbers for the conversion of benzene to phenol.
Co-reporter:Marit Wagner, ;Burkhard Ziemer
European Journal of Inorganic Chemistry 2008 Volume 2008( Issue 25) pp:3970-3976
Publication Date(Web):
DOI:10.1002/ejic.200800314
Abstract
Preparation routes for several sodium salts of diphenylbis(pyrazolyl)borates with bulky substituents are described. For a potential modelling of α-keto-acid-dependent nonheme iron enzymes, mononuclear iron compounds containing those bis(pyrazolyl)borates as ligands were synthesised and structurally characterised. The iron centres are four-coordinate (in one case only three-coordinate) thus offering room for the binding of exogenous ligands. However, reactivity studies with carboxylates and O2 have revealed nonuniform behaviour, and the identification of a side product in one case, combined with subsequent ESI-MS studies, demonstrated that a pronounced sensitivity of the B–N bonds towards oxo nucleophiles could be a possible reason.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2008)
Co-reporter:Marcus Brandt, Christian Limberg
Inorganica Chimica Acta 2008 Volume 361(Issue 2) pp:436-448
Publication Date(Web):15 January 2008
DOI:10.1016/j.ica.2007.03.019
The light-induced (546 nm) reaction of MnO3Cl with allene has been investigated in low-temperature argon matrices at 11 K. IR spectroscopic studies in combination with isotopic enrichment experiments (18O, D) and DFT calculations (B3LYP/LanL2DZ) allowed the identification of (O)2MnCl(OCCH2CH2) (1), and (O)2MnCl(H2COCCH2) (2) as the products. Possible ways for their formation are first of all discussed qualitatively in the context of the literature available, and then quantitatively with the background of DFT data (B3LYP/6-311G(d)) obtained for starting materials, products, transition states and intermediates. The most reasonable interpretation involves two-state reactivity.The light-induced reaction of MnO3Cl with allene in low-temperature argon matrices at 11 K leads to the products (O)2MnCl(OCCH2CH2) (1), and (O)2MnCl(H2COCCH2) (2). DFT studies revealed that 2 can be formed on the singlet surface via attack at the central or terminal carbon atoms, while a spin-crossover proceeds the formation of 1.
Co-reporter:Inke Siewert Dr.;Serhiy Demeshko Dr.;Elke Hoppe Dr.
Chemistry - A European Journal 2008 Volume 14( Issue 30) pp:9377-9388
Publication Date(Web):
DOI:10.1002/chem.200800955
Abstract
A ligand that offers two parallel malonate binding sites linked by a xanthene backbone, namely, Xanthmal2−, has been utilised to synthesise dinuclear FeII complex [Fe2(Xanthmal)2] (1). The reactivity of 1 in contact with O2 was investigated at −40 °C and room temperature. After activation of O2 through interaction with both iron centres the ligand is oxidised: at the Cα position monooxygenation and peroxide formation occur, partially accompanied by CC bond cleavage to yield α-keto ester groups. To reveal mechanistic details investigations concerning 1) peroxide decomposition, 2) the reactivity of a corresponding mononuclear complex, 3) the influence of monooxygenation of the ligand on the reactivity and 4) product formation in dependence on time were carried out. The results can be explained by postulating formation of high-valent Fe intermediates and ligand-to-metal electron transfer, and the mechanistic scheme derived includes several steps that mimic the (suggested) functioning of non-heme iron enzymes. In agreement with this proposal, ligand oxidation can also be performed catalytically. Furthermore, we show that via a competitive route [(Xanthmal)2Fe2O] (2) is formed, which is unreactive towards O2 and thus is a dead end with respect to ligand oxidation. Both compounds 1 and 2 were fully characterised, and their properties are discussed.
Co-reporter:Inke Siewert
Angewandte Chemie International Edition 2008 Volume 47( Issue 41) pp:7953-7956
Publication Date(Web):
DOI:10.1002/anie.200802955
Co-reporter:Inke Siewert
Angewandte Chemie 2008 Volume 120( Issue 41) pp:8071-8074
Publication Date(Web):
DOI:10.1002/ange.200802955
Co-reporter:Christian Limberg
European Journal of Inorganic Chemistry 2007 Volume 2007(Issue 21) pp:
Publication Date(Web):29 JUN 2007
DOI:10.1002/ejic.200700413
There are several reasons that make the combination of calixarenes with oxovanadium moieties attractive. Calixarenes are capable of simulating the surfaces of oxide bulk materials and supports, and on the other hand, supported vanadium oxides are an important class of heterogeneous catalysts, whose reactivity is still discussed controversially with respect to the effective mechanisms. Oxovanadium calixarene complexes can thus represent molecular models whose investigation in the homogeneous phase can shed some light into surface processes. This review describes the recent achievements in the synthesis of oxovanadium calixarene, thiacalixarene, and oxacalixarene complexes and the investigation of their reactivity. It has been shown that they represent active catalysts for the oxidation of alcohols and the polymerization/copolymerization of olefins.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2007)
Co-reporter:Elke Hoppe Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 24) pp:
Publication Date(Web):13 JUN 2007
DOI:10.1002/chem.200700354
With the aim of modeling reactive moieties and relevant intermediates on the surfaces of vanadium oxide based catalysts during oxygenation/dehydrogenation of organic substrates, mono- and dinuclear vanadium oxo complexes of doubly deprotonated p-tert-butylated tetrathiacalix[4]arene (H4TC) have been synthesized and characterized: PPh4[(H2TC)VOCl2] (1) and (PPh4)2[{(H2TC)V(O)(μ-O)}2] (2). According to the NMR spectra of the dissolved complexes they both retain the structures adopted in the crystalline state, as revealed by single-crystal X-ray crystallography. Compounds 1 and 2 were tested as catalysts for the oxidation of alcohols with O2 at 80 °C. Both 1 and 2 efficiently catalyze the oxidation of benzyl alcohol, crotyl alcohol, 1-phenyl-1-propanol, and fluorenol, and in most cases dinuclear complex 2 is more active than mononuclear complex 1. Moreover, the two thiacalixarene complexes 1 and 2 are in many instances more active than oxovanadium(V) complexes containing “classical” calixarene ligands tested previously. Complexes 1 and 2 also show significant activity in the oxidation of dihydroanthracene. Further investigations led to the conclusion that 1 acts as precatalyst that is converted to the active species PPh4[(TC)VO] (3) at 80 °C by double intramolecular HCl elimination. For complex 2, the results of mechanistic investigations indicated that the oxidation chemistry takes place at the bridging oxo ligands and that the two vanadium centers cooperate during the process. The intermediate (PPh4)2[{H2TCV(O)}2(μ-OH)(μ-OC13H9)] (4) was isolated and characterized, also with respect to its reactivity, and the results afforded a mechanistic proposal for a reasonable catalytic cycle. The implications which these findings gathered in solution may have for oxidation mechanisms on the surfaces of V-based heterogeneous catalysts are discussed.
Co-reporter:Robert M. Meudtner Dipl.-Chem.;Marc Ostermeier Dipl.-Chem.;Richard Goddard Dr.;Stefan Hecht Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 35) pp:
Publication Date(Web):19 OCT 2007
DOI:10.1002/chem.200701240
Click chemistry has been utilized to access 2,6-bis(1-aryl-1,2,3-triazol-4-yl)pyridines (BTPs) as versatile extended heteroaromatic building blocks for their exploitation in supramolecular chemistry, in particular foldamer and ligand design. In addition to their high-yielding synthesis using CuI-catalyzed Huisgen-type 1,3-dipolar cycloaddition reactions the formed triazole moieties constitute an integral part of the BTP framework and encode both its pronounced conformational preferences as well as its chelating ability. A diverse set of symmetrical and non-symmetrical BTPs carrying electron-donating and -withdrawing substituents at both terminal aryl and the central pyridine moieties has efficiently been synthesized and could furthermore readily be postfunctionalized with amphiphilic side chains and porphyrin chromophores. In both solution and solid state, the BTP scaffold adopts a highly conserved horseshoe-like anti–anti conformation. Upon protonation or metal coordination, the BTP scaffold switches to the chelating syn–syn conformation. Iron and europium complexes have been prepared, successfully characterized by single-crystal X-ray diffraction analysis, and investigated with regard to their spin state and luminescent properties. The extended heteroaromatic BTP scaffold should prove useful for the design of responsive foldamer backbones and the preparation of new magnetic and emissive materials.
Co-reporter:Marc Ostermeier Dr.;Burkhard Ziemer Dr.;Venugopal Karunakaran
Angewandte Chemie International Edition 2007 Volume 46(Issue 28) pp:
Publication Date(Web):1 JUN 2007
DOI:10.1002/anie.200700389
When having the blues isn't so bad: While normal coordination chemistry is observed when FeCl3 and the potential ligand 1 (see scheme) are mixed in dichloromethane, redox chemistry occurs in acetonitrile, which involves CH activation and CN bond formation, and finally leads to a water-soluble blue fluorophore 2.
Co-reporter:Marc Ostermeier Dr.;Burkhard Ziemer Dr.;Venugopal Karunakaran
Angewandte Chemie 2007 Volume 119(Issue 28) pp:
Publication Date(Web):1 JUN 2007
DOI:10.1002/ange.200700389
Blaumachen einmal anders: Der potenzielle Ligand 1 koordiniert in Dichlormethan in erwarteter Weise an FeCl3. Dagegen treten bei der Reaktion in Acetonitril Redoxprozesse auf, die C-H-Aktivierungen und C-N-Verknüpfungen zur Folge haben und schließlich zum wasserlöslichen blauen Fluorophor 2 führen.
Co-reporter:Stefan Roggan, Christian Limberg, Marcus Brandt, Burkhard Ziemer
Journal of Organometallic Chemistry 2005 Volume 690(Issue 23) pp:5282-5289
Publication Date(Web):15 November 2005
DOI:10.1016/j.jorganchem.2005.04.059
The reaction between Ph3BiBr2 and wet [NBu4]2[MoO4] leads to a white solid that according to Klemperer and Liu analyses as [NBu4]2[BiPh3(MoO4)2] · 3H2O (1aq). Working under strictly anhydrous conditions allowed us the isolation of the solvate-free complex [NBu4]2[BiPh3(MoO4)2] (1), which in contrast to 1aq could also be characterised by means of single crystal X-ray diffraction. The results reveal a structure with a BiV ion being surrounded by three phenyl substituents and two molybdate units. Remarkably the resulting two MoVI–O–BiV linkages are linear and according to a DFT investigation this is due to a predominantly ionic interaction between the O and Bi atoms. Moreover a novel MoVI–O–BiIII complex, NBu4[{Cp∗Mo(O)2-μ-O-}2(Bi(o-tolyl)2)] (2), has been prepared via reaction of the coordination polymer [(Cp∗Mo(O)2)-μ-O-(Bi(o-tolyl)2)]n with [NBu4][Cp∗MoO3] and the crystal structure of 2 has been investigated. According to DFT results the character of the bonds within the bent Mo–O–Bi unit is described most appropriately as covalent. The structure of 2 is discussed also with respect to corresponding Mo–O–Bi moieties occurring in bismuthmolybdate catalysts, for which it could represent a molecular structural model.Covalent Mo–O–Bi moities are found in the novel organometallic complex depicted, while structural analysis in combination with DFT calculations proved the interactions between molybdate anions and a bimuth(V) centre in a published compound to be of an electrostatic nature. Such complexes are interesting as molecular models for bismuthmolybdate catalysts.
Co-reporter:Stefan Roggan Dr.;Burkhard Ziemer Dr.
Angewandte Chemie 2005 Volume 117(Issue 33) pp:
Publication Date(Web):20 JUL 2005
DOI:10.1002/ange.200500375
Eine Brücke wird endlich gebaut: Das Umhüllen mit „weichen“ organischen Liganden ist der Schlüssel zur Isolierung von Komplexen mit „harten“ MoVI-O-BiV- und MoVI-O-BiIII-Einheiten (siehe Beispiel), die als molekulare Modellverbindungen für wichtige Spezies auf den Oberflächen von Heterogenkatalysatoren aufgefasst werden können.
Co-reporter:Christian Limberg Dr.
Angewandte Chemie 2005 Volume 117(Issue 38) pp:
Publication Date(Web):20 SEP 2005
DOI:10.1002/ange.200502936
Co-reporter:Stefan Roggan Dr.;Burkhard Ziemer Dr.
Angewandte Chemie International Edition 2005 Volume 44(Issue 33) pp:
Publication Date(Web):20 JUL 2005
DOI:10.1002/anie.200500375
Bridging the gap: Encapsulation by “soft” organic ligands is the key to the isolation of complexes containing “hard” MoVI-O-BiV and MoVI-O-BiIII units (see example), which might be regarded as molecular models for pivotal species on the surfaces of heterogeneous catalysts.
Co-reporter:Stefan Roggan;Gregor Schnakenburg Dr.;Steffen Shöfner Dr.;Hans Pritzkow Dr.;Burkhard Ziemer Dr.
Chemistry - A European Journal 2005 Volume 11(Issue 1) pp:
Publication Date(Web):17 NOV 2004
DOI:10.1002/chem.200400836
The reaction of molybdocenedihydride with two equivalents of [Bi(OtBu)3] proceeds via alcohol elimination and provides the compound [Cp2Mo{Bi(OtBu)2}2] (1), which contains two MoBi metal bonds, in good yields. If the two reagents are employed in a 1:1 ratio continuative condensation reactions occur. These initially lead to [{Cp2Mo}2{μ-Bi(OtBu)}2] (2), which, however, is very unstable in solution and decomposes via additional alcohol elimination: Complex-induced proximity effects facilitate the cleavage of CH bonds within the cyclopentadienyl ligands by the residual alkoxide ligands, so that spontaneously two further equivalents of alcohol are released, thereby yielding two isomeric compounds 3 and 4 with Cp ligands bridging MoBi metal bonds: The first isomer (3) contains two μ2-η5:η1-C5H4 ligands, the second isomer (4) contains one bridging μ3-η5:η1:η1-C5H3 ligand. The binding of these ligands to molybdenum and bismuth atoms at the same time is made possible through “bent bonds” between the bismuth and certain carbon centres. These unusual bonding situations were analysed by means of calculations based on density functional theory (DFT), the atoms in molecules (AIM) theory, natural bond order (NBO) considerations and the electron localisation function (ELF). According to the results the bonds can be understood in terms of carbanionic centres interacting with bismuth cations (i.e. closed-shell interactions). The formation of these bonds and the thermodynamics/kinetics involved on going from 2 to 3 and 4 were also studied by theoretical methods, so that the product formation is rationalised. The crystal structures of all four new compounds were determined. These structures but also the properties and mechanisms of formation are discussed against the background of the corresponding results obtained while studying the system [MeCp2MoH2]/[Bi(OtBu)3].
Co-reporter:Christian Limberg
Angewandte Chemie International Edition 2005 44(40) pp:6440
Publication Date(Web):
DOI:10.1002/anie.200590135
Co-reporter:Christian Limberg
Angewandte Chemie International Edition 2005 44(38) pp:6102-6104
Publication Date(Web):
DOI:10.1002/anie.200502936
Co-reporter:Cristina Wippert Rodrigues, Christian Limberg and Hans Pritzkow
Chemical Communications 2004 (Issue 23) pp:2734-2735
Publication Date(Web):15 Oct 2004
DOI:10.1039/B407475K
[Mo(η3-C3H5)(CO)2(bipy*)Cl] undergoes trigonal twist rearrangements in solution, so that three isomers are coexisting. It was used as a starting material leading to a dinuclear complex containing a hydrogen-bonded network of H2O and crown-ether molecules between two Mo(η3-C3H5)(CO)2(bipy*) moieties.
Co-reporter:Cristina Wippert Rodrigues;Hans Pritzkow
European Journal of Inorganic Chemistry 2004 Volume 2004(Issue 18) pp:
Publication Date(Web):12 JUL 2004
DOI:10.1002/ejic.200400209
In order to generate 2-propenylphenolate, PO−, as a potential ligand for oxomolybdenum compounds, the lithiation of the parent phenol derivative, POH, was investigated. Treatment of POH with BuLi in THF yielded a tetramer of the corresponding alkoxide “POLi” with a heterocubane structure in which the coordination spheres of the Li centres are completed by THF molecules, i.e. {Li(THF)[μ3-OC6H4(CH=CH−CH3)-2]}4 (1). Compound 1 proved to be rather unreactive and unsuitable as a starting material for the introduction of the corresponding aryloxide ligands into the coordination sphere of molybdenum by salt metathesis. The latter did occur, however, when “POLi” was synthesised in the presence of 12-crown-4 and subsequently treated with molybdenyl chloride. This led to an unidentified aryloxide compound as well as to a hitherto unknown chlorooxomolybdate anion in the complexes [Li(12-crown-4)][MoO2Cl3(THF)] (2) and [{Li(12-crown-4)}2Cl][MoO2Cl3(THF)] (2′) with an octahedral coordination sphere at the molybdenum centre in each case. An Mo(PO) complex was finally isolated for the first time when MoO2(OtBu)2 was employed as a starting material and treated with “POLi” prepared in situ in the presence of the crown ether. This led to the isolation of [Li(12-crown-4)2][MoO2(PO)3] (3), which represents a rare example of a five-coordinate dioxomolybdenum(VI) complex as well as being the first example among these complexes with a square-pyramidal ligand arrangement. Comparison of 2 with 3 suggests that the low coordination number of the 14-electron compound 3 originates from the bulkiness of the PO− ligands. The crystal structures of all three compounds 1−3 are discussed. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2004)
Co-reporter:Stefan Roggan Dr.;Burkhard Ziemer Dr.;Marcus Brt
Angewandte Chemie 2004 Volume 116(Issue 21) pp:
Publication Date(Web):12 MAY 2004
DOI:10.1002/ange.200353576
„Bananenbindungen“ zwischen Bi- und C-Atomen kennzeichnen 2, das sich durch intramolekulare Alkoholeliminierung aus 1 bildet. C-H-Aktivierungen, „complex induced proximity effects“, sind entscheidend für diese ungewöhnlichen Eliminierungen.
Co-reporter:Stefan Roggan Dr.;Burkhard Ziemer Dr.;Marcus Brt
Angewandte Chemie International Edition 2004 Volume 43(Issue 21) pp:
Publication Date(Web):12 MAY 2004
DOI:10.1002/anie.200353576
“Banana bonds” between Bi and C are a feature of 2, which forms in the course of intramolecular alcohol eliminations from 1. CH activations through complex-induced proximity effects are responsible for these unexpected eliminations.
Co-reporter:Christian Limberg Dr.
Angewandte Chemie International Edition 2003 Volume 42(Issue 48) pp:
Publication Date(Web):10 DEC 2003
DOI:10.1002/anie.200300578
The oxo-functionalization of organic substrates with the aid of metal oxo moieties is of fundamental importance not only in nature but also in academic and industrial research. Nevertheless the corresponding reaction mechanisms remain among the most enigmatic in chemistry and few of them are understood in detail. Recent research efforts have resulted in significantly improved information: In the cases of many oxygenation reactions evidence has been provided for the occurrence radical intermediates, even though the high selectivity observed suggests to a different mechanism. Examples stem from various areas of chemistry and include processes involving molecular metal oxo complexes, gas-phase and matrix-isolated species, metalloenzymes, and solid-state oxide surfaces. This Review treats this seemingly wide variety of systems with the aim of providing an overview of common reactivity patterns and principles, as well as open problems.
Co-reporter:Christian Limberg Dr.
Angewandte Chemie 2003 Volume 115(Issue 48) pp:
Publication Date(Web):10 DEC 2003
DOI:10.1002/ange.200300578
Metallvermittelte Oxofunktionalisierungen organischer Substrate sind von fundamentaler Bedeutung – in der Natur genauso wie in Hochschul- und Industrielaboratorien. Ungeachtet dessen sind die Mechanismen dieser Reaktionen rätselhaft geblieben, und nur wenige Prozesse werden im Detail verstanden. Dank intensiver Forschung konnte der Wissensstand in den letzten Jahren aber erheblich verbessert werden: So wurde gezeigt, dass viele Oxygenierungen über Radikalintermediate verlaufen – mitunter auch dann, wenn eine vergleichsweise hohe Selektivität Gegenteiliges vermuten lässt. Das zeigen Beispiele aus den unterschiedlichsten Bereichen: von molekularen Metalloxokomplexen, gasförmigen und matrixisolierten Spezies über das Reaktionsverhalten von Metalloenzymen bis hin zu Prozessen an Festkörperoberflächen. Der vorliegende Aufsatz beleuchtet die Vielfalt dieser Systeme und vermittelt einen Überblick über allgemein gültige Reaktionsmuster und Prinzipien sowie einige noch ungelöste Probleme.
Co-reporter:D. Pinkert, M. Keck, S. Ghassemi Tabrizi, C. Herwig, F. Beckmann, B. Braun-Cula, M. Kaupp and C. Limberg
Chemical Communications 2017 - vol. 53(Issue 57) pp:NaN8084-8084
Publication Date(Web):2017/06/26
DOI:10.1039/C7CC04670G
The famous α-Fe active sites in Fe-zeolites have recently been revealed to correspond to mononuclear high-spin iron(II) centres in square planar coordination environments. Here we report a first iron siloxide complex which represents a faithful structural and spectroscopic model of such sites. Notably, also an allogon with a distorted structure exists and could be crystallised.
Co-reporter:M. Sallmann, B. Braun and C. Limberg
Chemical Communications 2015 - vol. 51(Issue 31) pp:NaN6787-6787
Publication Date(Web):2015/03/19
DOI:10.1039/C5CC01083G
A novel complex TpMe,PhFe(SCH2CH2NH2) has been synthesized as a speculative model for ADO. Indeed its reaction with O2 led to the dioxygenation of the S atom and thus to hypotaurine. This finding may allow us to draw conclusions on the constitution of the ADO active site, whose structure is still unknown.
Co-reporter:Rafael Schiwon, Katharina Klingan, Holger Dau and Christian Limberg
Chemical Communications 2014 - vol. 50(Issue 1) pp:NaN102-102
Publication Date(Web):2013/11/11
DOI:10.1039/C3CC46629A
Modification of the Co-oxo cores of cobalt-polyoxometalate water oxidation catalysts is detectable by X-ray absorption spectroscopy (XAS) as demonstrated by comparison of Na10[Co4(H2O)2(PW9O34)2] (1) and Na17[((Co(H2O))Co2PW9O34)2(PW6O26)] (2). XAS reveals the integrity of 1 uncompromised by oxidant-driven water oxidation, which proceeds without formation of catalytic cobalt oxide.
Co-reporter:Bettina Horn, Christian Limberg, Christian Herwig and Beatrice Braun
Chemical Communications 2013 - vol. 49(Issue 93) pp:NaN10925-10925
Publication Date(Web):2013/10/17
DOI:10.1039/C3CC45407J
The β-diketiminato nickel(I) complex K2[LtBuNiI(N22−)NiILtBu] reacts with CO2via reductive disproportionation to form CO and CO32− containing products, whereas after employment of the NiI precursor [LtBuNiI(N2)NiILtBu] reductive coupling of CO2 was observed giving an oxalate bridged dinickel(II) complex. The addition of KC8 to the carbonate and oxalate compounds formed leads to the regeneration of the initial NiI complexes in an N2 atmosphere, thus closing synthetic cycles.
Co-reporter:Bettina Horn, Christian Limberg, Christian Herwig, Michael Feist and Stefan Mebs
Chemical Communications 2012 - vol. 48(Issue 66) pp:NaN8245-8245
Publication Date(Web):2012/07/12
DOI:10.1039/C2CC33846G
Reaction of a nickel(0) carbonyl complex, K2[LtBuNiCO]2, with N2O generates a cyclic carbonate compound composed of six [NiII(CO3)K]+ units. The same product can also be obtained using O2 as the oxidant in a solid-state/gas reaction. These conversions represent unique examples of a nickel-bound CO oxidation by N2O and O2, respectively.
Co-reporter:Peter Haack, Christian Limberg, Thomas Tietz and Ramona Metzinger
Chemical Communications 2011 - vol. 47(Issue 22) pp:NaN6376-6376
Publication Date(Web):2011/05/06
DOI:10.1039/C1CC11518A
The first structural characterisation of a copper–carbondisulfide complex revealed a hitherto unknown binding mode for CS2: it interacts with two metal centres (CuI) simultaneously via both CS π bonds. DFT calculations showed that complex formation occurs mainly due to a donation of electron density from the copper centres into the CS π* orbitals.
Co-reporter:Christina Knispel, Christian Limberg and Carolin Tschersich
Chemical Communications 2011 - vol. 47(Issue 38) pp:NaN10796-10796
Publication Date(Web):2011/09/05
DOI:10.1039/C1CC14616E
Reaction of [Cp2MoH2] with bismuth allyloxide, [Bi{OCH(CH3)CHCH2}3], gave rise to an extended octanuclear complex wherein two cyclic Mo2Bi2 units composed of four Mo–Bi bonds are linked by a Bi–Bi bond. The fact that the construction of such an assembly could be accomplished only in the case of a monomethylation of the parent allyl residue demonstrates a subtle substituent effect.
Co-reporter:Henrike Gehring, Ramona Metzinger, Beatrice Braun, Christian Herwig, Sjoerd Harder, Kallol Ray and Christian Limberg
Dalton Transactions 2016 - vol. 45(Issue 7) pp:NaN2996-2996
Publication Date(Web):2016/01/13
DOI:10.1039/C5DT04266F
After lithiation of PYR-H2 (PYR = [(NC(Me)C(H)C(Me)NC6H3(iPr)2)2(C5H3N)]2−) – the precursor of an expanded β-diketiminato ligand system with two binding pockets – with KN(TMS)2 the reaction of the resulting potassium salt with FeBr2 led to a dinuclear iron(II) bromide complex [(PYR)Fe(μ-Br)2Fe] (1). Through treatment with KHBEt3 the bromide ligands could be replaced by hydrides to yield [PYR)Fe2(μ-H)2] (2), a distorted analogue of known β-diketiminato iron hydride complexes, as evidenced by NMR, Mößbauer and X-ray absorption spectroscopy, as well as by its reactivity: for instance, 2 reacts with the proton source lutidinium triflate via protonation of the hydride ligands to form an iron(II) product [(PYR)Fe2(OTf)2] (4), while CO2 inserts into the Fe–H bonds generating the formate complex [(PYR)Fe2(μ-HCOO)2] (5); in the presence of traces of water partial hydrolysis occurs so that [(PYR)Fe2(μ-OH)(μ-HCOO)] (6) is isolated. Altogether, the iron(II) chemistry supported by the PYR2− ligand is distinctly different from the one of nickel(II), where both, the arrangement of the two binding pockets and the additional pyridyl donor led to diverging features as compared with the corresponding system based on the parent β-diketiminato ligand.
Co-reporter:Jan P. Falkenhagen, Christian Limberg, Serhiy Demeshko, Sebastian Horn, Michael Haumann, Beatrice Braun and Stefan Mebs
Dalton Transactions 2014 - vol. 43(Issue 2) pp:NaN816-816
Publication Date(Web):2013/10/23
DOI:10.1039/C3DT52349G
The reaction between [(TPA)Fe(MeCN)2](OTf)2 and [nBu4N](Cp*MoO3) yields the novel tetranuclear complex [(TPA)Fe(μ-Cp*MoO3)]2(OTf)2, 1, with a rectangular [Mo–O–Fe–O–]2 core containing high-spin iron(II) centres. 1 proved to be an efficient initiator/(pre)catalyst for the autoxidation of cis-cyclooctene with O2 to give cyclooctene epoxide. To test, which features of 1 are essential in this regard, analogues with zinc(II) and cobalt(II) central atoms, namely [(TPA)Zn(Cp*MoO3)](OTf), 3, and [(TPA)Co(Cp*MoO3)](OTf), 4, were prepared, which proved to be inactive. The precursor compounds of 1, [(TPA)Fe(MeCN)2](OTf)2 and [nBu4N](Cp*MoO3) as well as Cp2*Mo2O5, were found to be inactive, too. Reactivity studies in the absence of cyclooctene revealed that 1 reacts both with O2 and PhIO via loss of the Cp* ligands to give the triflate salt 2 of the known cation [((TPA)Fe)2(μ-O)(μ-MoO4)]2+. The cobalt analogue 4 reacts with O2 in a different way yielding [((TPA)Co)2(μ-Mo2O8)](OTf)2, 5, featuring a Mo2O84− structural unit which is novel in coordination chemistry. The compound [(TPA)Fe(μ-MoO4)]2, 6, being related to 1, but lacking Cp* ligands failed to trigger autoxidation of cyclooctene. However, initiation of autoxidation by Cp* radicals was excluded via experiments including thermal dissociation of Cp2*.
Co-reporter:Christian Ohde, Marcus Brandt, Christian Limberg, Jens Döbler, Burckhard Ziemer and Joachim Sauer
Dalton Transactions 2008(Issue 3) pp:NaN331-331
Publication Date(Web):2007/12/10
DOI:10.1039/B714933F
Inspired by surface species proposed to occur on heterogeneous catalysts novel oxovanadium(V) silsesquioxanes were synthesised. Reaction of a T8-silsequioxane containing two geminal OH groups with OV(OiPr)3 led to a dinuclear compound where the geminal disiloxide functions of two silsesquioxanes are bridging two OV(OiPr) moieties (2). Formation of 2 shows that—in contrast to proposals made for silica surfaces—in molecular chemistry a bidentate coordination of geminal siloxides to one vanadium centre is not favourable. With the background that species being doubly anchored to a support have been suggested to play active roles on V2O5/SiO2catalysts an anionic complex has been prepared where a divalent dioxovanadium unit replaces one Si corner of a (RSiO1.5)8 cube (a Si–OH function remains pending) (3). 3 has been intensely investigated by vibrational spectroscopy, and to support assignments not only of the ν(VO) bands but also of the ν(V–O–Si) bands, whose positions are of interest in the area of heterogeneous catalysis, isotopic enrichment studies and DFT calculations have been performed. The corresponding investigations were aided by the synthesis and analysis of a silylated derivative of 3, 4. Moreover, with regard to their potential as structural and spectroscopic models all complexes were characterised by single crystal X-ray diffraction. Finally, 2 and 3 were tested as potential catalysts for the photooxidation of cyclohexane and benzene with O2. While 2 shows a slightly higher activity than vanadylacetylacetonate, 3 leads to significantly increased turnover numbers for the conversion of benzene to phenol.
Co-reporter:Stefan Pfirrmann, Christian Limberg and Burkhard Ziemer
Dalton Transactions 2008(Issue 47) pp:NaN6691-6691
Publication Date(Web):2008/10/24
DOI:10.1039/B816136B
The preparation of a novel dinuclear nickel(II) hydride complex and its reactivity that often leads to nickel(I) compounds is described.