Co-reporter:Wen-Ching Chen, Wei-Chih Shih, Titel Jurca, Lili Zhao, Diego M. Andrada, Chun-Jung Peng, Chun-Chi Chang, Shu-kai Liu, Yi-Ping Wang, Yuh-Sheng Wen, Glenn P. A. Yap, Chao-Ping Hsu, Gernot Frenking, and Tiow-Gan Ong
Journal of the American Chemical Society September 13, 2017 Volume 139(Issue 36) pp:12830-12830
Publication Date(Web):August 16, 2017
DOI:10.1021/jacs.7b08031
An investigation of carbodicarbenes, the less explored member of the carbenic complex/ligand family has yielded unexpected electronic features and concomitant reactivity. Observed 1,2-addition of E–H bonds (E = B, C, Si) across the carbone central carbon and that of the flanking N-heterocyclic carbene (NHC) fragment, combined with single-crystal X-ray studies of a model Pd complex strongly suggests a significant level of π-accepting ability at the central carbon of the NHC moiety. This feature is atypical of classic NHCs, which are strong σ-donors, with only nominal π-accepting ability. The unanticipated π-acidity is critical for engendering carbodicarbenes with reactivity more commonly observed with frustrated Lewis pairs (FLPs) rather than the more closely related NHCs and cyclic (alkyl)(amino)carbenes (CAACs).
Co-reporter:Wen-Ching Chen, Wei-Chih Shih, Titel Jurca, Lili Zhao, Diego M. Andrada, Chun-Jung Peng, Chun-Chi Chang, Shu-kai Liu, Yi-Ping Wang, Yuh-Sheng Wen, Glenn P. A. Yap, Chao-Ping Hsu, Gernot Frenking, and Tiow-Gan Ong
Journal of the American Chemical Society September 13, 2017 Volume 139(Issue 36) pp:12830-12830
Publication Date(Web):August 16, 2017
DOI:10.1021/jacs.7b08031
An investigation of carbodicarbenes, the less explored member of the carbenic complex/ligand family has yielded unexpected electronic features and concomitant reactivity. Observed 1,2-addition of E–H bonds (E = B, C, Si) across the carbone central carbon and that of the flanking N-heterocyclic carbene (NHC) fragment, combined with single-crystal X-ray studies of a model Pd complex strongly suggests a significant level of π-accepting ability at the central carbon of the NHC moiety. This feature is atypical of classic NHCs, which are strong σ-donors, with only nominal π-accepting ability. The unanticipated π-acidity is critical for engendering carbodicarbenes with reactivity more commonly observed with frustrated Lewis pairs (FLPs) rather than the more closely related NHCs and cyclic (alkyl)(amino)carbenes (CAACs).
Co-reporter:Subrata Kundu, Prinson P. Samuel, Soumen Sinhababu, Anna V. Luebben, Birger Dittrich, Diego M. Andrada, Gernot Frenking, A. Claudia Stückl, Brigitte Schwederski, Alexa Paretzki, Wolfgang Kaim, and Herbert W. Roesky
Journal of the American Chemical Society August 16, 2017 Volume 139(Issue 32) pp:11028-11028
Publication Date(Web):July 26, 2017
DOI:10.1021/jacs.7b06610
The cyclic alkyl(amino) carbene (cAAC) stabilized biradicals of composition (cAAC)2SiH2 (1), (cAAC)SiMe2-SiMe2(cAAC) (2), and (cAAC)SiMeCl-SiMeCl(cAAC) (3) have been isolated as molecular species. All the compounds are stable at room temperature for more than 6 months under inert conditions in the solid state. All radical species were fully characterized by single-crystal X-ray structure analysis and EPR spectroscopy. Furthermore, the structure and bonding of compounds 1–3 have been investigated by theoretical methods. Compound 1 contains the SiH2 moiety and this is the first instance, where we have isolated 1 without an acceptor molecule.
Co-reporter:Jiaye Jin, Lili Zhao, Xiaonan Wu, Wei Li, Yuhong Liu, Diego M. Andrada, Mingfei Zhou, and Gernot Frenking
The Journal of Physical Chemistry A April 20, 2017 Volume 121(Issue 15) pp:2903-2903
Publication Date(Web):March 27, 2017
DOI:10.1021/acs.jpca.7b00739
The carbon suboxide cation C3O2+ and the protonated carbon suboxide HC3O2+/DC3O2+ were produced in the gas phase. The vibrational spectra were measured via infrared photodissociation spectroscopy of their argon- or CO-tagged complexes. Spectroscopic evidence combined with state-of-the-art quantum chemical calculations indicate that both cations have a bent C2v symmetry and can be designated as dicarbonyls of a carbon cation and methylidyne cation, respectively.
Co-reporter:Dr. Zhongshu Li;Dr. Xiaodan Chen;Dr. Diego M. Andrada; Dr. Gernot Frenking;Dr. Zoltán Benkö;Yaqi Li;Assoc. Dr. Jeffrey R. Harmer; Dr. Cheng-Yong Su; Dr. Hansjörg Grützmacher
Angewandte Chemie 2017 Volume 129(Issue 21) pp:5838-5843
Publication Date(Web):2017/05/15
DOI:10.1002/ange.201612247
AbstractCarbon phosphides, CnPm, may have highly promising electronic, optical, and mechanical properties, but they are experimentally almost unexplored materials. Phosphaheteroallenes stabilized by N-heterocyclic carbenes undergo a one-electron reduction to yield compounds of the type (L)2C2P2 with diverse structures. The use of imidazolylidenes as ligands L give complexes with a central four-membered ring C2P2, while more electrophilic cyclic diamidocarbenes (DAC) give a compound with an acyclic π-conjugated CP−PC unit. Cyclic C2P2 compounds are best described as non-Kekulé molecules that are stabilized by coordination to the NHC ligands NHC(C2P2)NHC. These species can be easily oxidized to give stable radical cations [(NHC)2C2P2]+.. The remarkably stable molecules with an acylic C2P2 core are best described with electron-sharing bonds (DAC)=C=P−P=C=(DAC).
Co-reporter:Dr. Simon J. Bonyhady;Dr. Nicole Holzmann; Gernot Frenking;Dr. Andreas Stasch; Cameron Jones
Angewandte Chemie 2017 Volume 129(Issue 29) pp:8647-8651
Publication Date(Web):2017/07/10
DOI:10.1002/ange.201610601
AbstractThe first example of a well-defined binary, low-oxidation-state aluminum hydride species that is stable at ambient temperature, namely the dianion in [{(DepNacnac)Mg}2(μ-H)]2[H3Al-AlH3] (DepNacnac=[(DepNCMe)2CH]−, Dep=2,6-diethylphenyl), has been prepared via a magnesium(I) reduction of the alanate complex, (DepNacnac)Mg(μ-H)3AlH(NEt3). An X-ray crystallographic analysis has shown the compound to be a contact ion complex, which computational studies have revealed to be the source of the stability of the aluminum(II) dianion.
Co-reporter:Dr. Zhongshu Li;Dr. Xiaodan Chen;Dr. Diego M. Andrada; Dr. Gernot Frenking;Dr. Zoltán Benkö;Yaqi Li;Assoc. Dr. Jeffrey R. Harmer; Dr. Cheng-Yong Su; Dr. Hansjörg Grützmacher
Angewandte Chemie International Edition 2017 Volume 56(Issue 21) pp:5744-5749
Publication Date(Web):2017/05/15
DOI:10.1002/anie.201612247
AbstractCarbon phosphides, CnPm, may have highly promising electronic, optical, and mechanical properties, but they are experimentally almost unexplored materials. Phosphaheteroallenes stabilized by N-heterocyclic carbenes undergo a one-electron reduction to yield compounds of the type (L)2C2P2 with diverse structures. The use of imidazolylidenes as ligands L give complexes with a central four-membered ring C2P2, while more electrophilic cyclic diamidocarbenes (DAC) give a compound with an acyclic π-conjugated CP−PC unit. Cyclic C2P2 compounds are best described as non-Kekulé molecules that are stabilized by coordination to the NHC ligands NHC(C2P2)NHC. These species can be easily oxidized to give stable radical cations [(NHC)2C2P2]+.. The remarkably stable molecules with an acylic C2P2 core are best described with electron-sharing bonds (DAC)=C=P−P=C=(DAC).
Co-reporter:Martin M. Juckel;Jamie Hicks;Dandan Jiang;Lili Zhao;Cameron Jones
Chemical Communications 2017 vol. 53(Issue 94) pp:12692-12695
Publication Date(Web):2017/11/23
DOI:10.1039/C7CC08430G
The first example of a stable zincagermylene, :Ge(TBoN)(ZnL*) (TBoN = N(SiMe3){B(DipNCH)2}, Dip = C6H3Pri2-2,6; L* = –N{C6H2[C(H)Ph2]2Me-2,6,4}(SiPri3)) is prepared and shown to have unprecedented reactivity for a germylene, with respect to the activation of dihydrogen. Computational analyses point towards this being partially derived from the electron releasing nature of the amido–zinc fragment, which leads to a narrowing of the HOMO–LUMO gap in the compound.
Co-reporter:Bin Li;Subrata Kundu;Hongping Zhu;Helena Keil;Regine Herbst-Irmer;Dietmar Stalke;Diego M. Andrada;Herbert W. Roesky
Chemical Communications 2017 vol. 53(Issue 17) pp:2543-2546
Publication Date(Web):2017/02/23
DOI:10.1039/C7CC00325K
The reaction of LAl: (L = HC[C(Me)N(Ar)]2, Ar = 2,6-iPr2C6H3) and cAAC:→AlX3 (X = Cl, I) (cAAC = cyclic alkyl amino carbene) results in strikingly asymmetric Al(II)–Al(II) compounds LAl(X)–Al(X)2–cAAC [X = Cl (2a); I (2b)]. In these dialuminum(II) compounds the two Al atoms bear different ligand environments. For a detailed insight into the structures, theoretical calculations were carried out.
Co-reporter:Rajendra S. Ghadwal;Dennis Rottschäfer;Diego M. Andrada;Christian J. Schürmann;Hans-Georg Stammler
Dalton Transactions 2017 vol. 46(Issue 24) pp:7791-7799
Publication Date(Web):2017/06/20
DOI:10.1039/C7DT01199G
The synthesis and characterization of the N-heterocyclic carbene (NHC) stabilized dichlorosilylene Group 6 metal complexes {(IPr)SiCl2}W(CO)5 (3-W), {(IPr)SiCl2}2Cr(CO)4 (4-Cr), and {(IPr)SiCl2}2W(CO)4 (4-W) (IPr = 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene) are reported. Treatment of 3-W with CsOH in the presence of IPr leads to the formation of an abnormal-NHC (aNHC) metal complex (aIPrH)W(CO)5 (6-W) (aIPrH = 1,3-bis(2,6-diisopropylphenyl)imidazol-4-ylidene), unveiling an unprecedented normal-to-abnormal transformation route of an NHC. DFT calculations support the proposed mechanism that involves CsOH mediated deprotonation of the IPr-backbone of 3-W to yield a ditopic carbanionic-NHC (dcNHC) complex 5a-W. Subsequent 1,4-migration of the W(CO)5 moiety and hydrolysis of the unmasked SiCl2 rationalize the formation of 6-W. The desired H2O molecule is generated in the initial step on deprotonation of IPr with CsOH. In contrast to the literature precedents, the calculations indicate that the abnormal complex 6-W is 13.5 kcal mol−1 thermodynamically higher in energy than the normal counterpart (IPr)W(CO)5 (8-W). Interestingly, as the aNHC-compounds reported so far are more stable than their normal counterparts, this finding showcases an opposite trend. Moreover, reaction pathways to the synthesized and related complexes have been investigated by DFT calculations.
Co-reporter:Subrata Kundu;Prinson P. Samuel;Anna Luebben;Diego M. Andrada;Birger Dittrich;Herbert W. Roesky
Dalton Transactions 2017 vol. 46(Issue 24) pp:7947-7952
Publication Date(Web):2017/06/20
DOI:10.1039/C7DT01796K
The cyclic alkyl(amino) carbene (cAAC) [:C{N-C6H3(2,6-IPr2)}(CMe2)2CH2] stabilized MeGeGeMe has been isolated in the molecular form with composition (cAAC)MeGe–GeMe(cAAC) (1) at room temperature. Compound 1 was synthesized from the reduction of MeGeCl3 using three equivalents of KC8 in the presence of one equivalent of cAAC. The corresponding silicon compound (cAAC)MeSi–SiMe(cAAC) (2) was also prepared. Compounds 1 and 2 are the first examples of REER compounds (E = Ge, Si) carrying the smallest organic group. Furthermore the structures of compounds 1 and 2 have been investigated by using theoretical methods. The theoretical analysis of the structure of 1 is in agreement with the formation as an unprecedented carbene stabilized bis-germylene whereas compound 2 can be equally described as carbene stabilized bis-silylene with coordinate bonds as with classical double bonds of a 2,3-disila-1,3-butadiene. The compounds were also characterized by X-ray crystallography.
Co-reporter:Zhuang Wu;Jian Xu;Qifan Liu;Xuelin Dong;Dingqing Li;Nicole Holzmann;Tarek Trabelsi;Joseph S. Francisco;Xiaoqing Zeng
Physical Chemistry Chemical Physics 2017 vol. 19(Issue 25) pp:16713-16720
Publication Date(Web):2017/06/28
DOI:10.1039/C7CP02774E
A biologically relevant reactive sulfur species (RSS), the hypothiocyanite radical OSCN, is generated in the gas phase through flash vacuum pyrolysis (FVP) of trifluoromethyl sulfinyl cyanide CF3S(O)CN at ca. 1000 K. Upon UV light irradiation (365 nm), OSCN rearranges to novel isomers OSNC and SOCN, and further visible light irradiation (400 ± 20 nm) leads to reverse isomerization. The identification of OSCN, OSNC, and SOCN in cryogenic matrices (Ar and N2, 2.8 K) with IR spectroscopy is supported by quantum chemical calculations up to the CCSD(T)-F12/VTZ-F12 level. The potential energy surface for the interconversion of OSCN isomers and their bonding properties are computationally explored by using the CCSD(T)-F12/VTZ-F12 and EDA–NOCV methods, respectively.
Co-reporter:Ranajit Saha;Sudip Pan;Pratim K. Chattaraj;Gabriel Merino
Physical Chemistry Chemical Physics 2017 vol. 19(Issue 3) pp:2286-2293
Publication Date(Web):2017/01/18
DOI:10.1039/C6CP06824C
A coupled-cluster study is performed on CO bound BeY complexes (Y = O, CO3, SO4, NH, NCN, and NBO) to understand the effect of attached ligands (Y) on the CO binding ability and C–O stretching frequency (νCO). Herein, we report that BeNCN has the highest CO binding ability (via both C- and O-side binding) among the studied neutral Be-based clusters, whereas OCBeSO4 has the highest νCO among the neutral carbonyls. The nature and extent of shift in νCO compared to free CO are explained in terms of change in polarization in the bonding orbitals of CO and relative contribution from OC→BeY or CO→BeY σ-donation, and OC←BeY or CO←BeY π-back-donation. The largest blue-shift in OCBeSO4 and the largest red-shift in COBeNH are consequences of the smallest OC←BeSO4 π-back-donation and the largest CO←BeNH π-back-donation, respectively.
Co-reporter:Lili Zhao, Markus Hermann, Nicole Holzmann, Gernot Frenking
Coordination Chemistry Reviews 2017 Volume 344(Volume 344) pp:
Publication Date(Web):1 August 2017
DOI:10.1016/j.ccr.2017.03.026
•The relevance of dative bonding in main group chemistry is discussed in the light of the progress which was made in the last decade.•The focus of the work lies on the analysis of the bonding situation using modern methods of quantum chemistry.The progress in experimental and theoretical research in the area of main group compounds that possess dative bonds, which was made in the last decade, is summarized. The focus lies on the information about the bonding situation that is available from the analysis of the electronic structure using modern methods of quantum chemistry and the implication on experimental work. The account describes complexes EL2 and E2L2 of group 13–15 atoms E where a single- or diatomic centre is stabilized by donor ligands L. The work is complemented by didactical excursions which demonstrate the strength of the dichotomous interpretation of chemical bonding in terms of dative or electron-sharing interactions. It is important to distinguish between the description of a bonding situation and the explanation of the molecular electronic structure, which provides the fundament for the pictorial representation of the bonds.
Co-reporter:Dr. Simon J. Bonyhady;Dr. Nicole Holzmann; Gernot Frenking;Dr. Andreas Stasch; Cameron Jones
Angewandte Chemie International Edition 2017 Volume 56(Issue 29) pp:8527-8531
Publication Date(Web):2017/07/10
DOI:10.1002/anie.201610601
AbstractThe first example of a well-defined binary, low-oxidation-state aluminum hydride species that is stable at ambient temperature, namely the dianion in [{(DepNacnac)Mg}2(μ-H)]2[H3Al-AlH3] (DepNacnac=[(DepNCMe)2CH]−, Dep=2,6-diethylphenyl), has been prepared via a magnesium(I) reduction of the alanate complex, (DepNacnac)Mg(μ-H)3AlH(NEt3). An X-ray crystallographic analysis has shown the compound to be a contact ion complex, which computational studies have revealed to be the source of the stability of the aluminum(II) dianion.
Co-reporter:Gernot Frenking, Markus Hermann, Diego M. Andrada and Nicole Holzmann
Chemical Society Reviews 2016 vol. 45(Issue 4) pp:1129-1144
Publication Date(Web):27 Jan 2016
DOI:10.1039/C5CS00815H
A summary of theoretical and experimental work in the area of low-coordinated compounds of boron and group-14 atoms C–Sn in the last decade is presented. The focus of the account lies on molecules EL2, E2L2 and E3L3, which possess dative bonds between one, two or three atoms E and σ-donor ligands L that stabilize the atoms E through L→E donor–acceptor interactions. The interplay between theory and experiment provides detailed insight into the bonding situation of the molecules, which serves as guideline for the synthesis of molecules that possess unusual bonding motifs.
Co-reporter:Chandrajeet Mohapatra, Subrata Kundu, Alexander N. Paesch, Regine Herbst-Irmer, Dietmar Stalke, Diego M. Andrada, Gernot Frenking, and Herbert W. Roesky
Journal of the American Chemical Society 2016 Volume 138(Issue 33) pp:10429-10432
Publication Date(Web):August 5, 2016
DOI:10.1021/jacs.6b07361
The cyclic alkyl(amino) carbene stabilized Si2H2 has been isolated in the molecular form of composition (Me-cAAC:)2Si2H2 (1) and (Cy-cAAC:)2Si2H2 (2) at room temperature. Compounds 1 and 2 were synthesized from the reduction of HSiCl3 using 3 equiv of KC8 in the presence of 1 equiv of Me-cAAC: and Cy-cAAC:, respectively. These are the first molecular examples of Si2H2 characterized by single crystal X-ray structural analysis. Moreover, electrospray ionization mass spectrometry and 1H as well as 29Si NMR data are reported. Furthermore, the structure of compound 1 has been investigated by theoretical methods. The theoretical analysis of 1 explains equally well its structure with coordinate bonds as with classical double bonds of a 2,3-disila-1,3-butadiene.
Co-reporter:Hui Wang;Christian Bethke;Markus Hermann;Seema Agarwal
Macromolecular Chemistry and Physics 2016 Volume 217( Issue 16) pp:1834-1841
Publication Date(Web):
DOI:10.1002/macp.201600154
Co-reporter:Markus Hermann ; Gernot Frenking
Chemistry - A European Journal 2016 Volume 22( Issue 12) pp:4100-4108
Publication Date(Web):
DOI:10.1002/chem.201503762
Abstract
Quantum chemical calculations using the complete active space of the valence orbitals have been carried out for HnCCHn (n=0–3) and N2. The quadratic force constants and the stretching potentials of HnCCHn have been calculated at the CASSCF/cc-pVTZ level. The bond dissociation energies of the C−C bonds of C2 and HC≡CH were computed using explicitly correlated CASPT2-F12/cc-pVTZ-F12 wave functions. The bond dissociation energies and the force constants suggest that C2 has a weaker C−C bond than acetylene. The analysis of the CASSCF wavefunctions in conjunction with the effective bond orders of the multiple bonds shows that there are four bonding components in C2, while there are only three in acetylene and in N2. The bonding components in C2 consist of two weakly bonding σ bonds and two electron-sharing π bonds. The bonding situation in C2 can be described with the σ bonds in Be2 that are enforced by two π bonds. There is no single Lewis structure that adequately depicts the bonding situation in C2. The assignment of quadruple bonding in C2 is misleading, because the bond is weaker than the triple bond in HC≡CH.
Co-reporter:Terrance J. Hadlington, Markus Hermann, Gernot Frenking and Cameron Jones
Chemical Science 2015 vol. 6(Issue 12) pp:7249-7257
Publication Date(Web):22 Sep 2015
DOI:10.1039/C5SC03376D
Reactions of the solution stable, two-coordinate hydrido-tetrylenes, :E(H)(L†) (E = Ge or Sn; L† = –N(Ar†)(SiPri3); Ar† = C6H2{C(H)Ph2}2Pri-2,6,4), with a variety of unactivated cyclic and acyclic alkenes, and one internal alkyne, lead to the rapid and regiospecific hydrometallation of the unsaturated substrate at ambient temperature. The products of the reactions, [L†E(C2H4R)] (E = Ge or Sn, R = H, Ph or But), [L†E{CH(CH2)3(CH2)n}] (E = Ge, n = 1, 2 or 3; E = Sn, n = 1) and [L†E{C(Ph)C(H)(Me)}], include the first structurally characterised examples of two-coordinate amido/alkyl germylenes and stannylenes. The cycloalkene hydrometallation reactions are cleanly reversible under ambient conditions, a process which computational and experimental van't Hoff analyses suggest proceeds via β-hydride elimination from the metal coordinated cycloalkyl ligand. Similarly, the reactions of :Ge(H)(L†) with 1,5-cyclooctadiene and 2-methyl-2-butene, both likely proceed via β-hydride elimination processes, leading to the clean isomerisation of the alkene involved, and its subsequent hydrogermylation, to give [L†Ge(2-cyclooctenyl)] and [L†Ge{C2H4C(H)Me2}], respectively. Reactions of [L†GeEt] and [L†Ge(C5H9)] with the protic reagents, HCl, NH3 and EtOH, lead to oxidative addition to the germanium(II) centre, and formation of the stable chiral germanium(IV) complexes, [L†Ge(C5H9)(H)Cl] and [L†Ge(Et)(H)R] (R = NH2 or OEt). In contrast, related reactions between [L†SnEt] and ButOH or TEMPOH (TEMP = 2,2,6,6-tetramethylpiperidinyl) proceed via ethane elimination, affording the tin(II) products, [L†SnR] (R = OBut or OTEMP). In addition, the oxidation of [L†Ge(C6H11)] and [L†Sn(C2H4But)] with O2 yields the oxo-bridged metal(IV) dimers, [{L†(C6H11)Ge(μ-O)}2] and [{L†(ButC2H4)Sn(μ-O)}2], respectively.
Co-reporter:Nicole Holzmann, Markus Hermann and Gernot Frenking
Chemical Science 2015 vol. 6(Issue 7) pp:4089-4094
Publication Date(Web):01 Jun 2015
DOI:10.1039/C5SC01504A
Quantum chemical calculations of the compound B2(NHCMe)2 and a thorough examination of the electronic structure with an energy decomposition analysis provide strong evidence for the appearance of boron–boron triple bond character. This holds for the model compound and for the isolated diboryne B2(NHCR)2 of Braunschweig which has an even slightly shorter B–B bond. The bonding situation in the molecule is best described in terms of NHCMe→B2←NHCMe donor–acceptor interactions and concomitant π-backdonation NHCMe←B2→NHCMe which weakens the B–B bond, but the essential features of a triple bond are preserved. An appropriate formula which depicts both interactions is the sketch NHCMe⇄BB⇄NHCMe. Calculations of the stretching force constants FBB which take molecules that have genuine single, double and triple bonds as references suggest that the effective bond order of B2(NHCMe)2 has the value of 2.34. The suggestion by Köppe and Schnöckel that the strength of the boron–boron bond in B2(NHCH)2 is only between a single and a double bond is repudiated. It misleadingly takes the force constant FBB of OBBO as the reference value for a B–B single bond which ignores π bonding contributions. The alleged similarity between the B–O bonds in OBBO and the B–C bonds in B2(NHCMe)2 is a mistaken application of the principle of isolable relationship.
Co-reporter:Kerstin Freitag; Hung Banh; Christian Gemel; Paul Jerabek; Rüdiger W. Seidel; Gernot Frenking;Roland A. Fischer
Inorganic Chemistry 2015 Volume 54(Issue 1) pp:352-358
Publication Date(Web):December 17, 2014
DOI:10.1021/ic502532g
The synthesis and characterization of the cationic mixed metal Ga/Zn cluster [Zn2(GaCp*)6]2+ (1) is presented. The reaction of [Zn2Cp*2] with [Ga2Cp*][BAr4F] leads to the formation of the novel complex being the first example of a [Zn2]2+ core exclusively ligated by metalloid group-13 organyl-ligands. Compound 1 exhibits two different coordination modes: In the solid state, two of the six GaCp* ligands occupy bridging positions, whereas VT 1H NMR indicates the coexistence of a second isomer in solution featuring six terminal GaCp* ligands. Quantum chemical calculations have been carried out to assign the gallium and zinc positions; the bonding situation in 1 is characterized and the importance of dispersion forces is discussed.
Co-reporter: Gernot Frenking; Giovanni F. Caramori
Angewandte Chemie International Edition 2015 Volume 54( Issue 9) pp:2596-2599
Publication Date(Web):
DOI:10.1002/anie.201411374
Co-reporter:Dr. Kerstin Freitag;Dr. Christian Gemel;Dr. Paul Jerabek;Dr. Iris M. Oppel;Dr. Rüdiger W. Seidel;Dr. Gernot Frenking;Hung Banh;Katharina Dilchert;Dr. Rol A. Fischer
Angewandte Chemie 2015 Volume 127( Issue 14) pp:4445-4449
Publication Date(Web):
DOI:10.1002/ange.201410737
Abstract
Die dreieckigen Cluster [Zn3Cp*3]+ und [Zn2CuCp*3] wurden durch Addition der in situ generierten, elektrophilen und isolobalen Spezies [ZnCp*]+ und [CuCp*] an das Carmona-Reagens [Cp*ZnZnCp*] synthetisiert, wobei die Zn-Zn-Bindung des Letzteren erhalten blieb. Die Verwendung von nichtkoordinierenden fluorierten aromatischen Lösungsmitteln ist dabei essenziell. Die Bindungssituation der ligandenstabilisierten Cluster wurde mit quantenchemischen Methoden untersucht, wobei sich ein hohes Maß an σ-Aromatizität zeigte, ähnlich der Bindungssituation im Triwasserstoffkation [H3]+. Die neuen Spezies dienen als molekulare Bausteine für CunZnm-Cluster (Nanomessing), die auch im LIFDI-Massenspektrum der neuen Verbindungen detektiert werden können.
Co-reporter: Gernot Frenking; Giovanni F. Caramori
Angewandte Chemie 2015 Volume 127( Issue 9) pp:2632-2635
Publication Date(Web):
DOI:10.1002/ange.201411374
Co-reporter:Qingnan Zhang, Mohua Chen, Mingfei Zhou, Diego M. Andrada, and Gernot Frenking
The Journal of Physical Chemistry A 2015 Volume 119(Issue 11) pp:2543-2552
Publication Date(Web):October 16, 2014
DOI:10.1021/jp509006u
The novel neon complex NeBeCO3 has been prepared in a low-temperature neon matrix via codeposition of laser-evaporated beryllium atoms with O2 + CO/Ne. Doping by the heavier noble gas atoms argon, krypton and xenon yielded the associated adducts NgBeCO3 (Ng = Ar, Kr, Xe). The noble gas complexes have been identified via infrared spectroscopy. Quantum chemical calculations of NgBeCO3 and NgBeO (Ng = He, Ne, Ar, Kr, Xe) using ab initio methods and density functional theory show that the Ng–BeCO3 bonds are slightly longer and weaker than the Ng–BeO bonds. The energy decomposition analysis of the Ng–Be bonds suggests that the attractive interactions come mainly from the Ng → BeCO3 and Ng → BeO σ donation.
Co-reporter:Dr. Kerstin Freitag;Dr. Christian Gemel;Dr. Paul Jerabek;Dr. Iris M. Oppel;Dr. Rüdiger W. Seidel;Dr. Gernot Frenking;Hung Banh;Katharina Dilchert;Dr. Rol A. Fischer
Angewandte Chemie International Edition 2015 Volume 54( Issue 14) pp:4370-4374
Publication Date(Web):
DOI:10.1002/anie.201410737
Abstract
The triangular clusters [Zn3Cp*3]+ and [Zn2CuCp*3] were obtained by addition of the in situ generated, electrophilic, and isolobal species [ZnCp*]+ and [CuCp*] to Carmona’s compound, [Cp*ZnZnCp*], without splitting the ZnZn bond. The choice of non-coordinating fluoroaromatic solvents was crucial. The bonding situations of the all-hydrocarbon-ligand-protected clusters were investigated by quantum chemical calculations revealing a high degree of σ-aromaticity similar to the triatomic hydrogen ion [H3]+. The new species serve as molecular building units of CunZnm nanobrass clusters as indicated by LIFDI mass spectrometry.
Co-reporter: Mohua Chen;Qingnan Zhang; Mingfei Zhou;Dr. Diego M. Andrada; Gernot Frenking
Angewandte Chemie International Edition 2015 Volume 54( Issue 1) pp:124-128
Publication Date(Web):
DOI:10.1002/anie.201406264
Abstract
The complexes OCBeCO3 and COBeCO3 have been isolated in a low-temperature neon matrix. The more stable isomer OCBeCO3 has a very high CO stretching mode of 2263 cm−1, which is blue-shifted by 122 cm−1 with respect to free CO and 79 cm−1 higher than in OCBeO. Bonding analysis of the complexes shows that OCBeO has a stronger OCBeY bond than OCBeCO3 because it encounters stronger π backdonation. The isomers COBeCO3 and COBeO exhibit red-shifted CO stretching modes with respect to free CO. The inverse change of CO stretching frequency in OCBeY and COBeY is explained with the reversed polarization of the σ and π bonds in CO.
Co-reporter:Terrance J. Hadlington, Jiaye Li, Markus Hermann, Amelia Davey, Gernot Frenking, and Cameron Jones
Organometallics 2015 Volume 34(Issue 13) pp:3175-3185
Publication Date(Web):June 15, 2015
DOI:10.1021/acs.organomet.5b00206
Reactions of two sterically hindered amido-digermynes, L*GeGeL* (1; L* = −N(Ar*)(SiMe3); Ar* = C6H2Me{C(H)Ph2}2-4,2,6) and L†GeGeL† (2; L† = −N(Ar†)(SiPri3); Ar† = C6H2Pri{C(H)Ph2}2-4,2,6), with a variety of olefins and related molecules are investigated. These lead to the facile reduction, C–H activation, dehydrogenation, and/or cycloaddition of the unsaturated substrate. Specifically, reaction of L†GeGeL† with ethylene proceeds via a formal [2 + 2 + 2] cycloaddition to give the digermabicyclo[2.2.0]hexane L†Ge(μ-C2H4)2GeL† (3). In contrast, treating L†GeGeL† with norbornadiene proceeds via reductive insertion of one olefin moiety of the organic substrate into the Ge–Ge bond of 1, yielding the norbornenediyl-bridged bis(germylene) L†Ge(μ-C7H8)GeL† (4). Similarly, L*GeGeL* doubly reduces cyclooctatetraene (COT) to give the planar cyclooctateraenediyl inverse sandwich complex L*Ge(μ-η2,η2-COT)GeL* (5). An indication that this reaction occurs via an initial formal [2 + 2] cycloaddition intermediate comes from the reaction of L†GeGeL† with 1,5-cyclooctadiene (COD). This affords the [2 + 2] cycloaddition product L†Ge(COD)GeL† (6), which exists in solution in equilibrium with 2 and free COD. A computational study indicates that 6 readily dissociates, as the reaction that gave it is close to thermoneutral. Reaction of 1,3-cyclohexadiene (1,3-CHD) with L†GeGeL† yields the 1,4-bis(germylene) substituted cyclohex-2-enediyl L†Ge(μ-C6H8)GeL† (7), which is an isolated intermediate in the transfer hydrogenation, or C–H activation, reaction between L†GeGeL† and 1,3-CHD. Heating 7 gives benzene and the known digermene L†(H)GeGe(H)L†. Reactions of 1 or 2 with propyne, bis(trimethylsilyl)butadiyne, and azobenzene all lead to reductive insertion of the unsaturated substrate into the Ge–Ge bond of the digermyne and formation of L†Ge{μ-HC═C(Me)}GeL† (8), L*Ge{μ-(Me3Si)C═C(CCSiMe3)}GeL* (9), and L*Ge{μ-(Ph)NN(Ph)}GeL* (10), respectively. The reaction of 4-dimethylaminopyridine (DMAP) with L*GeGeL* gives the adduct complex L*(DMAP)GeGe(DMAP)L* (11). Taken as a whole, this study highlights both similarities and significant differences between the reactivities of the amido-digermynes 1 and 2 and those of their previously described terphenyl-substituted counterparts.
Co-reporter:Dayne C. Georgiou;Bradley D. Stringer;Dr. Conor F. Hogan;Dr. Peter J. Barnard;Dr. David J. D. Wilson;Dr. Nicole Holzmann;Dr. Gernot Frenking;Dr. Jason L. Dutton
Chemistry - A European Journal 2015 Volume 21( Issue 8) pp:3377-3386
Publication Date(Web):
DOI:10.1002/chem.201405416
Abstract
The attempted synthesis of NHC-stabilized dicarbon (NHCCCNHC) through deprotonation of a doubly protonated precursor ([NHCCHCHNHC]2+) is reported. Rather than deprotonation, a clean reduction to NHCCHCHNHC is observed with a variety of bases. The apparent resistance towards deprotonation to the target compound led to a reinvestigation of the electronic structure of NHCCCNHC, which showed that the highest occupied molecular orbital/lowest unoccupied molecular orbital (HOMO/LUMO) gap is likely too small to allow for isolation of this species. This is in contrast to the recent isolation of the cyclic alkylaminocarbene analogue (cAACCCcAAC), which has a large HOMO–LUMO gap. A detailed theoretical study illuminates the differences in electronic structures between these molecules, highlighting another case of the potential advantages of using cAAC rather than NHC as a ligand. The bonding analysis suggests that the dicarbon compounds are well represented in terms of donor–acceptor interactions LC2L (L=NHC, cAAC).
Co-reporter:Gernot Frenking, Ralf Tonner, Susanne Klein, Nozomi Takagi, Takayazu Shimizu, Andreas Krapp, Krishna K. Pandey and Pattiyil Parameswaran
Chemical Society Reviews 2014 vol. 43(Issue 14) pp:5106-5139
Publication Date(Web):11 Jun 2014
DOI:10.1039/C4CS00073K
Recent theoretical studies are reviewed which show that the naked group 14 atoms E = C–Pb in the singlet 1D state behave as bidentate Lewis acids that strongly bind two σ donor ligands L in the donor–acceptor complexes L→E←L. Tetrylones EL2 are divalent E(0) compounds which possess two lone pairs at E. The unique electronic structure of tetrylones (carbones, silylones, germylones, stannylones, plumbylones) clearly distinguishes them from tetrylenes ER2 (carbenes, silylenes, germylenes, stannylenes, plumbylenes) which have electron-sharing bonds R–E–R and only one lone pair at atom E. The different electronic structures of tetrylones and tetrylenes are revealed by charge- and energy decomposition analyses and they become obvious experimentally by a distinctively different chemical reactivity. The unusual structures and chemical behaviour of tetrylones EL2 can be understood in terms of the donor–acceptor interactions L→E←L. Tetrylones are potential donor ligands in main group compounds and transition metal complexes which are experimentally not yet known. The review also introduces theoretical studies of transition metal complexes [TM]–E which carry naked tetrele atoms E = C–Sn as ligands. The bonding analyses suggest that the group-14 atoms bind in the 3P reference state to the transition metal in a combination of σ and π∥ electron-sharing bonds TM–E and π⊥ backdonation TM→E. The unique bonding situation of the tetrele complexes [TM]–E makes them suitable ligands in adducts with Lewis acids. Theoretical studies of [TM]–E→W(CO)5 predict that such species may becomes synthesized.
Co-reporter:David S. Weinberger ; Nurul Amin SK ; Kartik Chandra Mondal ; Mohand Melaimi ; Guy Bertrand ; A. Claudia Stückl ; Herbert W. Roesky ; Birger Dittrich ; Serhiy Demeshko ; Brigitte Schwederski ; Wolfgang Kaim ; Paul Jerabek
Journal of the American Chemical Society 2014 Volume 136(Issue 17) pp:6235-6238
Publication Date(Web):April 14, 2014
DOI:10.1021/ja502521b
Two (cAAC)2Cu complexes, featuring a two-coordinate copper atom in the formal oxidation state zero, were prepared by reducing (Et2-cAAC)2Cu+I− with metallic sodium in THF, and by a one-pot synthesis using Me2-cAAC, Cu(II)Cl2, and KC8 in toluene in a molar ratio of 2:1:2, respectively. Both complexes are highly air and moisture sensitive but can be stored in the solid state for a month at room temperature. DFT calculations showed that in these complexes the copper center has a d10 electronic configuration and the unpaired electron is delocalized over two carbene carbon atoms. This was further confirmed by the EPR spectra, which exhibit multiple hyperfine lines due to the coupling of the unpaired electron with 63,65Cu isotopes, 14N, and 1H nuclei.
Co-reporter:Kartik Chandra Mondal ; Prinson P. Samuel ; Herbert W. Roesky ; Elena Carl ; Regine Herbst-Irmer ; Dietmar Stalke ; Brigitte Schwederski ; Wolfgang Kaim ; Liviu Ungur ; Liviu F. Chibotaru ; Markus Hermann
Journal of the American Chemical Society 2014 Volume 136(Issue 5) pp:1770-1773
Publication Date(Web):January 17, 2014
DOI:10.1021/ja4123285
(Me2-cAAC:)2Co2 (2, where Me2-cAAC: = cyclic alkyl amino carbene, :C(CH2)(CMe2)2N-2,6-iPr2C6H3)) was synthesized via the reduction of precursor (Me2-cAAC:CoII(μ-Cl)Cl)2 (1) with KC8. 2 contains two cobalt atoms in the formal oxidation state zero. Magnetic measurement revealed that 2 has a singlet spin ground state S = 0. The cyclic voltammogram of 2 exhibits both one-electron oxidation and reduction, indicating the possible synthesis of stable species containing 2•– and 2•+ ions. The latter was synthesized via reduction of 1 with required equivalents of KC8 and characterized as [(Me2-cAAC:)2Co2]•+OTf– (2•+OTf–). Electron paramagnetic resonance spectroscopy of 2•+ reveals the coupling of the electron spin with 2 equiv 59Co isotopes, leading to a (Co0.5)2 state. The experimental Co1–Co2 bond distances are 2.6550(6) and 2.4610(6) Å for 2 and 2•+OTf–, respectively. Theoretical investigation revealed that both 2 and 2•+OTf– possess a Co–Co bond with an average value of 2.585 Å. A slight increase of the Co–Co bond length in 2 is more likely to be caused by the strong π-accepting property of cAAC. 2•+ is only 0.8 kcal/mol higher in energy than the energy minimum. The shortening of the Co–Co bond of 2•+ is caused by intermolecular interactions.
Co-reporter:Sudipta Roy; Peter Stollberg; Regine Herbst-Irmer; Dietmar Stalke; Diego M. Andrada; Gernot Frenking;Herbert W. Roesky
Journal of the American Chemical Society 2014 Volume 137(Issue 1) pp:150-153
Publication Date(Web):December 24, 2014
DOI:10.1021/ja512089e
The unstable species dichlorosilylene was previously stabilized by carbene. The lone pair of electrons on the silicon atom of (carbene)SiCl2 can form a coordinate bond with metal–carbonyls. Herein we report that (carbene)SiCl2 can stabilize a phosphinidene (Ar–P, a carbone analogue) with the general formula carbene→SiCl2→P–Ar (carbene = cyclic alkyl(amino) carbene (cAAC; 2) and N-heterocyclic carbene (NHC; 3)). Compounds 2 and 3 are stable, isolable, and storable at 0 °C (2) to room temperature (3) under an inert atmosphere. The crystals of 2 and 3 are dark blue and red, respectively. The intense blue color of 2 arises due to the strong intramolecular charge transfer (ICT) transition from πSi═P→π*cAAC. The electronic structure and bonding of 2, 3 were studied by theoretical calculations. The HOMO of the molecule is located on the πSi═P bond, while the LUMO is located at the carbene moiety (cAAC or NHC). The dramatic change in color of these compounds from red (3, NHC) to blue (2, cAAC) is ascribed to the difference in energy of the LUMO within the carbenes (cAAC/NHC) due to a lower lying LUMO of cAAC.
Co-reporter:Fabian Dielmann ; Diego M. Andrada ; Gernot Frenking ;Guy Bertrand
Journal of the American Chemical Society 2014 Volume 136(Issue 10) pp:3800-3802
Publication Date(Web):February 22, 2014
DOI:10.1021/ja5007355
Transition metal complexes featuring a metal–nitrogen multiple bond have been widely studied due to their implication in dinitrogen fixation and catalytic nitrogen–carbon bond formation. Terminal copper– and silver–nitrene complexes have long been proposed to be the key intermediates in aziridination and amination reactions using azides as the nitrogen source. However, due to their high reactivity, these species have eluded isolation and spectroscopic characterization even at low temperatures. In this paper we report that a stable phosphinonitrene reacts with coinage metal trifluoromethanesulfonates, affording bridging and terminal copper– and silver–nitrene complexes, which are characterized by NMR spectroscopy and single crystal X-ray diffraction analysis.
Co-reporter:Paul Jerabek ; Herbert W. Roesky ; Guy Bertrand
Journal of the American Chemical Society 2014 Volume 136(Issue 49) pp:17123-17135
Publication Date(Web):November 13, 2014
DOI:10.1021/ja508887s
Quantum chemical calculations using density functional theory have been carried out for the cyclic (alkyl)(amino)carbene (cAAC) complexes of the group 11 atoms [TM(cAAC)2] (TM = Cu, Ag, Au) and their cations [TM(cAAC)2]+. The nature of the metal–ligand bonding was investigated with the charge and energy decomposition analysis EDA-NOCV. The calculations show that the TM–C bonds in the charged adducts [TM(cAAC)2]+ are significantly longer than in the neutral complexes [TM(cAAC)2], but the cations have much higher bond dissociation energies than the neutral molecules. The intrinsic interaction energies ΔEint in [TM(cAAC)2]+ take place between TM+ in the 1S electronic ground state and (cAAC)2. In contrast, the metal–ligand interactions in [TM(cAAC)2] involve the TM atoms in the excited 1P state yielding strong TM p(π) → (cAAC)2 π backdonation, which is absent in the cations. The calculations suggest that the cAAC ligands in [TM(cAAC)2] are stronger π acceptors than σ donors. The trends of the intrinsic interaction energies and the bond dissociation energies of the metal–ligand bonds in [TM(cAAC)2] and [TM(cAAC)2]+ give the order Au > Cu > Ag. Calculations at the nonrelativistic level give weaker TM–C bonds, particularly for the gold complexes. The trend for the bond strength in the neutral and charged adducts without relativistic effects becomes Cu > Ag > Au. The EDA-NOCV calculations suggest that the weaker bonds at the nonrelativistic level are mainly due to stronger Pauli repulsion and weaker orbital interactions. The NBO picture of the C–TM–C bonding situation does not correctly represent the nature of the metal–ligand interactions in [TM(cAAC)2].
Co-reporter:Terrance J. Hadlington ; Markus Hermann ; Gernot Frenking ;Cameron Jones
Journal of the American Chemical Society 2014 Volume 136(Issue 8) pp:3028-3031
Publication Date(Web):February 13, 2014
DOI:10.1021/ja5006477
This study details the first use of well-defined low-valent p-block metal hydrides as catalysts in organic synthesis. That is, the bulky, two-coordinate germanium(II) and tin(II) hydride complexes, L†(H)M: (M = Ge or Sn, L† = −N(Ar†)(SiPri3), Ar† = C6H2{C(H)Ph2}2Pri-2,6,4), are shown to act as efficient catalysts for the hydroboration (with HBpin, pin = pinacolato) of a variety of unactivated, and sometimes very bulky, carbonyl compounds. Catalyst loadings as low as 0.05 mol % are required to achieve quantitative conversions, with turnover frequencies in excess of 13 300 h–1 in some cases. This activity rivals that of currently available catalysts used for such reactions.
Co-reporter:Yan Li, Hongping Zhu, Diego M. Andrada, Gernot Frenking and Herbert W. Roesky
Chemical Communications 2014 vol. 50(Issue 35) pp:4628-4630
Publication Date(Web):10 Mar 2014
DOI:10.1039/C4CC00912F
An interesting aminosilanetrithiol RSi(SH)3 (R = N(SiMe3)-2,6-iPr2C6H3) has been prepared by the reaction of lithium aminosilanetrithiolate {RSi[SLi(THF)]3}2 with MeCOOH. Theoretical calculations indicate that the LP(N) → σ*(Si–S) and LP(S) → σ*(Si–S) electron donations remarkably contribute to the stabilization of the Si(SH)3 part of the molecule. RSi(SH)3 is the first example of a stable molecule containing three SH groups attached to one element.
Co-reporter:Markus Hermann, Cameron Jones, and Gernot Frenking
Inorganic Chemistry 2014 Volume 53(Issue 13) pp:6482-6490
Publication Date(Web):May 5, 2014
DOI:10.1021/ic500457q
The calculated reaction profiles using density functional theory at the BP86/TZVPP level for the reaction of small molecules with amidoditetrylynes R2N–EE–NR2 (E = Si, Ge, Sn) are discussed. Four projects are presented that feature the virtue of cooperation between theory and experiment. First, the calculated reaction paths for hydrogenation of the model systems (Me2N)EEL(NMe2) (E = Si, Ge, Sn), which possess E–E single bonds, are examined. The results for the germanium model systems are compared with hydrogenation of the real system L†GeGeL† where L† = NAr*(SiMe3) (Ar* = C6H2{C(H)Ph2}2Me-2,6,4). The second project introduced the multiply bonded amidodigermyne L††GeGeL††, which carries the extremely bulky substituents L†† = N(Ar††)(SiPri3), where Ar†† = C6H2{C(H)Ph2}2Pri-2,6,4. The theoretical reaction profile for dihydrogen addition to L††GeGeL†† is discussed. Hydrogenation gives L††(H)GeGe(H)L†† as the product, which is in equilibrium with the hydrido species Ge(H)L††. The latter germanium hydride and tin homologue Sn(H)L†† were found to be effective catalysts for hydroboration reactions, which is the topic of the third project. Finally, the calculated reaction course for the reduction of CO2 to CO with the amidodigermyne L†GeGeL† is discussed.
Co-reporter:Arik Puls;Paul Jerabek;Wataru Kurashige;Moritz Förster;Mariusz Molon;Dr. Timo Bollermann;Manuela Winter;Dr. Christian Gemel;Dr. Yuichi Negishi;Dr. Gernot Frenking;Dr. Rol A. Fischer
Angewandte Chemie 2014 Volume 126( Issue 17) pp:4415-4419
Publication Date(Web):
DOI:10.1002/ange.201310436
Abstract
Heterometalldotierte Goldcluster sind durch nasschemische Synthese schwer zugänglich, und mit Hauptgruppenmetallen oder frühen Übergangsmetallen dotierte Cluster sind rar. Die Verbindungen [M(AuPMe3)11(AuCl)]3+ (M=Pt, Pd, Ni) (1–3), [Ni(AuPPh3)(8–2n)(AuCl)3(AlCp*)n] (n=1, 2) (4, 5) und [Mo(AuPMe3)8(GaCl2)3(GaCl)]+ (6) wurden selektiv durch Transmetallierung von [M(M′Cp*)n] (M=Mo, E=Ga, n=6; M=Pt, Pd, Ni, M′=Ga, Al; n=4) mit [ClAuPR3] (R=Me, Ph) hergestellt und mit Einkristallröntgenbeugung sowie ESI-MS charakterisiert. Mithilfe von DFT-Rechnungen wurden die Bindungsverhältnisse analysiert. Die Transmetallierung ist eine wirkungsvolle Syntheseroute hin zu heterometalldotierten Goldclustern, deren Aufbau der 18-Valenzelektronenregel für das Zentralmetallatom gehorcht und die mit dem Superatomkonzept auf Grundlage des Jellium-Modells übereinstimmen.
Co-reporter:Alexer S. Ivanov;Dr. Alexer I. Boldyrev;Dr. Gernot Frenking
Chemistry - A European Journal 2014 Volume 20( Issue 9) pp:2431-2435
Publication Date(Web):
DOI:10.1002/chem.201304566
Abstract
A theoretical study of Li90P90, which possesses a circular double-helix structure that resembles the Watson–Crick DNA structure, is reported. This is a new bonding motif in inorganic chemistry. The calculations show that the molecule might become synthesized and that it could be a model for other inorganic species which possess a double-helix structure.
Co-reporter:Alexander S. Ivanov, Gernot Frenking, and Alexander I. Boldyrev
The Journal of Physical Chemistry A 2014 Volume 118(Issue 35) pp:7375-7384
Publication Date(Web):January 28, 2014
DOI:10.1021/jp4123997
Despite the confirmation of Cl––Cl– association in aqueous solution and crystalline state, there have been no reports about the existence of stable dichloride anion pair in the gas phase. In the current work we performed a systematic ab initio study of microsolvation of dichloride anion pair. The stepwise solvation mechanism observed for free gaseous [Cl2(H2O)n]2– (n = 2–10) clusters was found to be quite interesting. The lowest structure for dichloride hexahydrate closely resembles cubic water octamer W8 in which two water molecules in the corners of the cube are substituted by two chloride anions. We have also shown that Cl––Cl– pair may be completely stabilized by about 36 water molecules in the gas phase. Stabilization of the pair leads to the formation of cyclic H2O structures that bridge the Cl– ions. It has been predicted that the large clusters of [Cl2(H2O)36]2– and [Cl2(H2O)40]2– may exhibit properties analogous to bulk aqueous solutions, therefore they could become good molecular models for understanding complicated processes of solvation of Cl– in the bulk.
Co-reporter:Gerhard Hilt, Judith Janikowski, Martin Schwarzer, Olaf Burghaus, Dimitri Sakow, Martin Bröring, Marcel Drüschler, Benedikt Huber, Bernhard Roling, Klaus Harms, Gernot Frenking
Journal of Organometallic Chemistry 2014 749() pp: 219-223
Publication Date(Web):
DOI:10.1016/j.jorganchem.2013.09.020
Co-reporter:Dr. Jing Xu;Dr. Yi-Hong Ding;Dr. Diego M. Andrada;Dr. Gernot Frenking
Chemistry - A European Journal 2014 Volume 20( Issue 30) pp:9216-9220
Publication Date(Web):
DOI:10.1002/chem.201403252
Abstract
Quantum chemical calculations show that the N-heterocyclic carbene (NHC)-stabilized silavinylidene, NHCtBuCSiR2 is a strongly bonded complex, which has a linear arrangement of the donor and acceptor moieties. The molecule is the energetically lowest lying isomer of the NHC-stabilized R2CSi isomers and it is stable towards dimerization when R is a bulky substituent. The silavinylidene complex is the only species of the silylidene homologues NHCtBuEE′R2 (E, E′=C–Pb) which possesses a linear arrangement. The unusual bonding situation is explained in terms of donor–acceptor interactions between NHCtBu as σ donor and CSiR2 in the doubly excited singlet state 3a12b2 which leads to a significantly shorter CSiR2 bond compared with free CSiR2.
Co-reporter:Dr. Gernot Frenking
Angewandte Chemie International Edition 2014 Volume 53( Issue 24) pp:6040-6046
Publication Date(Web):
DOI:10.1002/anie.201311022
Co-reporter:Arik Puls;Paul Jerabek;Wataru Kurashige;Moritz Förster;Mariusz Molon;Dr. Timo Bollermann;Manuela Winter;Dr. Christian Gemel;Dr. Yuichi Negishi;Dr. Gernot Frenking;Dr. Rol A. Fischer
Angewandte Chemie International Edition 2014 Volume 53( Issue 17) pp:4327-4331
Publication Date(Web):
DOI:10.1002/anie.201310436
Abstract
Heterometal-doped gold clusters are poorly accessible through wet-chemical approaches and main-group-metal- or early-transition-metal-doped gold clusters are rare. Compounds [M(AuPMe3)11(AuCl)]3+ (M=Pt, Pd, Ni) (1–3), [Ni(AuPPh3)(8−2n)(AuCl)3(AlCp*)n] (n=1, 2) (4, 5), and [Mo(AuPMe3)8 (GaCl2)3(GaCl)]+ (6) were selectively obtained by the transmetalation of [M(M′Cp*)n] (M=Mo, E=Ga, n=6; M=Pt, Pd, Ni, M′=Ga, Al, n=4) with [ClAuPR3] (R=Me, Ph) and characterized by single-crystal X-ray diffraction and ESI mass spectrometry. DFT calculations were used to analyze the bonding situation. The transmetalation proved to be a powerful tool for the synthesis of heterometal-doped gold clusters with a design rule based on the 18 valence electron count for the central metal atom M and in agreement with the unified superatom concept based on the jellium model.
Co-reporter:Dr. Gernot Frenking
Angewandte Chemie 2014 Volume 126( Issue 24) pp:6152-6158
Publication Date(Web):
DOI:10.1002/ange.201311022
Co-reporter:Yuki Ito ; Vladimir Ya. Lee ; Heinz Gornitzka ; Catharina Goedecke ; Gernot Frenking ;Akira Sekiguchi
Journal of the American Chemical Society 2013 Volume 135(Issue 18) pp:6770-6773
Publication Date(Web):April 17, 2013
DOI:10.1021/ja401650q
In this contribution, we report a spirobis(pentagerma[1.1.1]propellane) derivative as a novel type of molecular architecture in cluster chemistry that features two spiro-fused [1.1.1]propellane units and represents a stable tetraradicaloid species. The crucial issue of the nature of the interaction between the germanium bridgeheads was probed computationally, revealing weak bonding interactions between the formally unpaired electrons.
Co-reporter:Amit Pratap Singh ; Prinson P. Samuel ; Herbert W. Roesky ; Martin C. Schwarzer ; Gernot Frenking ; Navdeep S. Sidhu ;Birger Dittrich
Journal of the American Chemical Society 2013 Volume 135(Issue 19) pp:7324-7329
Publication Date(Web):April 21, 2013
DOI:10.1021/ja402351x
Metal ions with radical centers in their coordination sphere are key participants in biological and catalytic processes. In the present study, we describe the synthesis of the cAAC:ZnCl2 adduct (1) using a cyclic alkylaminocarbene (cAAC) as donor ligand. Compound 1 was treated with 2 equiv of KC8 and LiB(sec-Bu)3H to yield a deep blue-colored dicarbene zinc compound (cAAC)2Zn (2) and the colorless hydrogenated zinc compound (cAACH)2Zn (3), respectively. Compounds 2 and 3 were well characterized by spectroscopic methods and single-crystal X-ray structural analysis. Density functional theory calculations were performed for 2 which indicate that this molecule possesses a singlet biradicaloid character. Moreover, we show the application of 2 in CO2 activation, which yields a zwitterionic cAAC·CO2 adduct.
Co-reporter:Chelladurai Ganesamoorthy, Sinah Loerke, Christian Gemel, Paul Jerabek, Manuela Winter, Gernot Frenking and Roland A. Fischer
Chemical Communications 2013 vol. 49(Issue 28) pp:2858-2860
Publication Date(Web):22 Jan 2013
DOI:10.1039/C3CC38584A
Compounds Cp*AlH2 (1) and Cp*2AlH (2) reductively eliminate Cp*H in benzene or toluene under reflux conditions to give Al(s) and AlCp*, respectively.
Co-reporter:Mehmet Ali Celik, Chandrakanta Dash, Venkata A. K. Adiraju, Animesh Das, Muhammed Yousufuddin, Gernot Frenking, and H. V. Rasika Dias
Inorganic Chemistry 2013 Volume 52(Issue 2) pp:729-742
Publication Date(Web):December 28, 2012
DOI:10.1021/ic301869v
N-Heterocyclic carbene ligand SIDipp (SIDipp = 1,3-bis(2,6-diisopropylphenyl)imidazolin-2-ylidene) and trimesitylphosphine ligand have been used in the synthesis of gold(I) cyanide, t-butylisocyanide, and cyclooctyne complexes (SIDipp)Au(CN) (3), (Mes3P)Au(CN) (4), [(Mes3P)2Au][Au(CN)2] (5), [(SIDipp)Au(CNtBu)][SbF6] ([6][SbF6]), [(SIDipp)Au(cyclooctyne)][SbF6] ([8][SbF6]), and [(Mes3P)Au(cyclooctyne)][SbF6] ([9][SbF6]). A detailed computational study has been carried out on these and the related gold(I) carbonyl adducts [(SIDipp)Au(CO)][SbF6] ([1][SbF6]), [(Mes3P)Au(CO)][SbF6] ([2][SbF6]), and [(Mes3P)Au(CNtBu)]+ ([7]+). X-ray crystal structures of 3, 5, [6][SbF6], [8][SbF6], and [9][SbF6] revealed that they feature linear gold sites. Experimental and computational data show that the changes in π-acid ligand on (SIDipp)Au+ from CO, CN–, CNtBu, cyclooctyne as in [1]+, 3, [6]+, and [8]+ did not lead to large changes in the Au–Ccarbene bond distances. A similar phenomenon was also observed in Au–P distance in complexes [2]+, 4, [7]+, and [9]+ bearing trimesitylphosphine. Computational data show that the Au–L bonds of “naked” [Au–L]+ or SIDipp and Mes3P supported [Au–L]+ (L = CO, CN–, CNtBu to cyclooctyne) have higher electrostatic character than covalent character. The Au←L σ-donation and Au→L π-back-donation contribute to the orbital term with the former being the dominant component, but the latter is not negligible. In the Au–CO adducts [1]+and [2]+, the cationic gold center causes the polarization of the C–O σ and π orbitals toward the carbon end making the coefficients at the two atoms more equal which is mainly responsible for the large blue shift in the CO stretching frequency. The SIDipp and Mes3P supported gold(I) complexes of cyanide and isocyanide also exhibit a significant blue shift in υ̅CN compared to that of the free ligands. Calculated results for Au(CO)Cl and Au(CF3)CO suggest that the experimentally observed blue shift in ν̅CO of these compounds may at least partly be caused by intermolecular forces.
Co-reporter:Mariusz Molon, Christian Gemel, Rüdiger W. Seidel, Paul Jerabek, Gernot Frenking, and Roland A. Fischer
Inorganic Chemistry 2013 Volume 52(Issue 12) pp:7152-7160
Publication Date(Web):May 23, 2013
DOI:10.1021/ic400741b
Organozinc (ZnR with R = Cp*, Me, Cl, Br) ligated transition metal (M) half-sandwich compounds of general formula [Cp*M(ZnR)5] (M = Fe, Ru) are presented in this work. The new compounds were obtained by treatment of various GaCp* ligated precursors with suitable amounts of ZnMe2 to exchange Ga against Zn. This exchange follows a strict Ga:Zn ratio of 1:2. Accordingly, a Ga/Zn mixed compound [{Cp*Ru(GaCp*)(ZnCp*)(ZnCl)2}2] can be obtained if the amount of ZnMe2 is reduced so that one GaCp* remains coordinated to the transition metal. All new compounds were characterized by elemental analysis, 1H and 13C NMR spectroscopy as well as by single crystal X-ray diffraction techniques, if applicable. The coordination polyhedra of [Cp*M(ZnR)5] can be derived from the pseudo homoleptic parent compound [Ru(ZnCp*)4(ZnMe)6], as emphasized by continuous shape measures analysis (CShM). Computational investigations at the density functional theory (DFT) level of theory were performed, revealing no significant attractive interaction of the zinc atoms and therefore these compounds are best described as classical complexes, rather than cluster compounds. The Ru-L bond strength follow the order Cp* > ZnCl > ZnMe > ZnCp*.
Co-reporter:Shannon A. Couchman, Nicole Holzmann, Gernot Frenking, David J. D. Wilson and Jason L. Dutton
Dalton Transactions 2013 vol. 42(Issue 32) pp:11375-11384
Publication Date(Web):02 Apr 2013
DOI:10.1039/C3DT50563D
A theoretical study of compounds containing Be in the +1 or 0 oxidation state has been carried out. The molecules considered containing Be in the +1 oxidation state are analogues of the important Mg(I)–Mg(I) dimer supported by the β-diketiminate ligand. The molecules in the 0 oxidation state are NHC supported compounds analogous to “molecular allotropes” which has recently become a topic of importance in p-block chemistry. In this case, our results demonstrate that the Be(0) complexes are far more stable than the analogous Mg(0) complexes, highlighting the opportunities afforded in Be chemistry, despite the challenges presented by the toxicity of Be compounds.
Co-reporter:Catharine Esterhuysen and Gernot Frenking
Dalton Transactions 2013 vol. 42(Issue 37) pp:13349-13356
Publication Date(Web):23 May 2013
DOI:10.1039/C3DT32872D
DFT calculations using BP86 in conjunction with the SVP and TZVPP basis sets as well as ab initio calculations at SCS-MP2 have been carried out for six dicoordinated carbon molecules CLL′ where L is a fluorenyl carbene while L′ is a phosphine PH3 (1) or PPh3 (2) or a carbene, i.e. NHCMe (3), benzannulated NHCMe (4), cycloheptatrienylidene (5) and benzannulated cycloheptatrienylidene (6). The complexes of these compounds with one and two AuCl moieties were also calculated. The monoaurated adducts of 1–4 have the AuCl fragment η1 coordinated to the central carbon atom. The complexes 5(AuCl) and 6(AuCl) have AuCl η2 bonded across a CC double bond. Three different bonding modes are found as energy minima for the diaurated species LL′C-(AuCl)2. The AuCl fragments are found to be either both coordinated η1, both coordinated η2 across double bonds, or a combination of the two. According to the electronic structure analysis of the free compounds, 1 and 2 might best be classified as carbenes, 3 and 4 as bent allenes while 5 and 6 are typical allenes. The complexation with AuCl reveals that 1–4 may exhibit chemical behaviour which is typical for carbones and thus, they may be termed “hidden carbones”. The AuCl complexes show that compounds 5 and 6 are classical allenes.
Co-reporter:Paul Jerabek, Gernot Frenking
International Journal of Mass Spectrometry 2013 Volumes 354–355() pp:342-345
Publication Date(Web):15 November 2013
DOI:10.1016/j.ijms.2013.06.030
Quantum chemical ab intio calculations at the MP2/6-311++G(2d2f)//MP2/6-311G(df) level have been carried out for the neutral and charged cyclobutane-1,2,3,4-tetrathione C4S4q species 3aq (q = −1, 0, +1) and for the tricyclic bis-1,2-dithiete isomers 3bq. The equilibrium geometries of 3aq possess a square-planar D4h symmetry where the CC bonds of neutral 3a become longer in 3a+ and shorter in 3a− while the CS distances become shorter in the cation 3a+ and longer in the anion 3a−. The bis-1,2-dithiete isomers 3bq are significantly higher in energy than the tetrathione forms 3aq. Neutral 3b which is 67.9 kcal/mol less stable than 3a possesses a trans-ladder arrangement of the four-membered rings with two short (1.390 Å) and two longer (1.515 Å) CC bonds. The cation 3b+ which is 57.4 kcal/mol less stable than 3a+ has a distorted boat-shaped arrangement of the four-membered rings where the CS bond lengths of the 1,2-dithete rings are clearly different. The anion 3b− which is 108.3 kcal/mol higher in energy than 3a− has a trans-ladder type structure like the neutral parent system 3b.
Co-reporter:Markus Hermann, Catharina Goedecke, Cameron Jones, and Gernot Frenking
Organometallics 2013 Volume 32(Issue 22) pp:6666-6673
Publication Date(Web):September 26, 2013
DOI:10.1021/om4007888
Quantum chemical calculations of the reaction profiles for addition of one and two H2 molecules to amido-substituted ditetrylynes have been carried using density functional theory at the BP86/def2-TZVPP//BP86/def2-TZVPP level of theory for the model systems L′EEL′ and BP86/def2-TZVPP//BP86/def-SVP for the real compounds. The hydrogenation of the digermyne LGeGeL (L = N(SiMe3)Ar*; Ar* = C6H2Me{C(H)Ph2}2-4,2,6) follows a stepwise reaction course. The addition of the first H2 gives the singly bridged species LGe(μ-H)GeHL, which rearranges with very low activation barriers to the symmetrically hydrogenated compound LHGeGeHL and to the most stable isomer LGeGe(H)2L, which is experimentally observed. The addition of the second H2 proceeds with a higher activation energy under rupture of the Ge–Ge bond, yielding LGeH and LGeH3 as reaction products. Energy calculations which consider dispersion interactions using Grimme’s D3 term suggest that the latter reaction is thermodynamically unfavorable. The second hydrogenation reaction LGeGe(H)2L → L(H)2GeGe(H)2L possesses an even higher activation barrier than the bond-breaking hydrogenation step. Further calculations which consider solvent effects change the theoretically predicted reaction profile very little. The calculations of the model system L′GeGeL′ (L′ = NMe2) give a very similar reaction profile. Calculations of the model disilyne and distannyne homologues L′SiSiL′ and L′SnSnL′ suggest that the reactivity of the amido-substituted ditetrylynes always has the order Si > Ge > Sn. The most stable product of the addition of one H2 to the distannyne L′SnSnL′ is the doubly bridged species L′Sn(μ-H)2SnL′, which has been experimentally observed when bulky groups are employed. Analysis of the H2–L′EEL′ interactions in the transition state for the addition of the first H2 with the EDA-NOCV method reveals that the HOMO–LUMO and LUMO–HOMO interactions have similar magnitudes.
Co-reporter:Adinarayana Doddi;Dr. Christian Gemel;Manuela Winter;Dr. Rol A Fischer;Catharina Goedecke;Dr. Henry S. Rzepa;Dr. Gernot Frenking
Angewandte Chemie International Edition 2013 Volume 52( Issue 1) pp:450-454
Publication Date(Web):
DOI:10.1002/anie.201204440
Co-reporter:Dr. Kartik Chra Mondal;Dr. Herbert W. Roesky;Martin C. Schwarzer;Dr. Gernot Frenking;Dr. Igor Tkach;Hilke Wolf;Daniel Kratzert;Dr. Regine Herbst-Irmer;Benedikt Niepötter;Dr. Dietmar Stalke
Angewandte Chemie International Edition 2013 Volume 52( Issue 6) pp:1801-1805
Publication Date(Web):
DOI:10.1002/anie.201204487
Co-reporter: Gernot Frenking;Markus Hermann
Angewandte Chemie International Edition 2013 Volume 52( Issue 23) pp:5922-5925
Publication Date(Web):
DOI:10.1002/anie.201301485
Co-reporter:Dr. Prinson P. Samuel;Dr. Kartik Chra Mondal;Dr. Herbert W. Roesky;Markus Hermann;Dr. Gernot Frenking;Dr. Serhiy Demeshko;Dr. Franc Meyer;Dr. A. Claudia Stückl;Jonathan H. Christian;Dr. Naresh S. Dalal;Dr. Liviu Ungur; Liviu F. Chibotaru;Dr. Kevin Pröpper;Dr. Alke Meents;Priv.-Doz.Dr. Birger Dittrich
Angewandte Chemie International Edition 2013 Volume 52( Issue 45) pp:11817-11821
Publication Date(Web):
DOI:10.1002/anie.201304642
Co-reporter:Dipl.-Chem. Nicole Holzmann;Dr. Andreas Stasch; Cameron Jones; Gernot Frenking
Chemistry - A European Journal 2013 Volume 19( Issue 20) pp:6467-6479
Publication Date(Web):
DOI:10.1002/chem.201203679
Abstract
Quantum chemical calculations using density functional theory at the BP86/TZ2P level have been carried out to determine the geometries and stabilities of Group 13 adducts [(PMe3)(EH3)] and [(PMe3)2(E2Hn)] (E=B–In; n=4, 2, 0). The optimized geometries exhibit, in most cases, similar features to those of related adducts [(NHCMe)(EH3)] and [(NHCMe)2(E2Hn)] with a few exceptions that can be explained by the different donor strengths of the ligands. The calculations show that the carbene ligand L=NHCMe (:C(NMeCH)2) is a significantly stronger donor than L=PMe3. The equilibrium geometries of [L(EH3)] possess, in all cases, a pyramidal structure, whereas the complexes [L2(E2H4)] always have an antiperiplanar arrangement of the ligands L. The phosphine ligands in [(PMe3)2(B2H2)], which has Cs symmetry, are in the same plane as the B2H2 moiety, whereas the heavier homologues [(PMe3)2(E2H2)] (E=Al, Ga, In) have Ci symmetry in which the ligands bind side-on to the E2H2 acceptor. This is in contrast to the [(NHCMe)2(E2H2)] adducts for which the NHCMe donor always binds in the same plane as E2H2 except for the indium complex [(NHCMe)2(In2H2)], which exhibits side-on bonding. The boron complexes [L2(B2)] (L=PMe3 and NHCMe) possess a linear arrangement of the LBBL moiety, which has a BB triple bond. The heavier homologues [L2(E2)] have antiperiplanar arrangements of the LEEL moieties, except for [(PMe3)2(In2)], which has a twisted structure in which the PInInP torsion angle is 123.0°. The structural features of the complexes [L(EH3)] and [L2(E2Hn)] can be explained in terms of donor–acceptor interactions between the donors L and the acceptors EH3 and E2Hn, which have been analyzed quantitatively by using the energy decomposition analysis (EDA) method. The calculations predict that the hydrogenation reaction of the dimeric magnesium(I) compound L′MgMgL′ with the complexes [L(EH3)] is energetically more favorable for L=PMe3 than for NHCMe.
Co-reporter:Terrance J. Hadlington;Markus Hermann;Jiaye Li;Dr. Gernot Frenking; Cameron Jones
Angewandte Chemie 2013 Volume 125( Issue 39) pp:10389-10393
Publication Date(Web):
DOI:10.1002/ange.201305689
Co-reporter:Animesh Das, Chandrakanta Dash, Mehmet Ali Celik, Muhammed Yousufuddin, Gernot Frenking, and H. V. Rasika Dias
Organometallics 2013 Volume 32(Issue 11) pp:3135-3144
Publication Date(Web):March 6, 2013
DOI:10.1021/om400073a
The tris(alkyne) copper complex [(cyclooctyne)3Cu][SbF6] has been synthesized using cyclooctyne and in situ generated CuSbF6. Tris(alkyne) silver complexes [(cyclooctyne)3Ag]+ involving weakly coordinating counterions such as [SbF6]− and [PF6]− have also been isolated in good yield using cyclooctyne and commercially available AgSbF6 and AgPF6. These coinage metal tris(alkyne) adducts have trigonal-planar metal sites. The alkyne carbon atoms and the metal site form distorted spoke-wheel (rather than upright trigonal-prismatic) structures in the solid state. In [(cyclooctyne)3Cu][SbF6], these distortions result in a propeller-like arrangement of alkynes. A cationic gold(I) complex having two alkynes has been prepared by a reaction of equimolar amounts of Au(cyclooctyne)2Cl and AgSbF6 in dichloromethane. The gold atom of [(cyclooctyne)2Au]+ coordinates to the cyclooctynes in a linear fashion, while the carbon atoms of the alkyne groups form a tetrahedron around gold(I). Optimized geometries of cationic [(cyclooctyne)3M]+, [(cyclooctyne)2M]+, and [(cyclooctyne)M]+ and neutral [(cyclooctyne)2MCl] and [(cyclooctyne)MCl] adducts (M = Cu, Ag, Au) using density functional theory (DFT) at the BP86/def2-TZVPP level of theory and a detailed analysis of metal–alkyne bonding interactions are also presented.
Co-reporter:Masoumeh Mousavi and Gernot Frenking
Organometallics 2013 Volume 32(Issue 6) pp:1743-1751
Publication Date(Web):March 11, 2013
DOI:10.1021/om301163w
Quantum chemical calculations using gradient corrected density functional theory at the BP86/def2-TZVPP level have been carried out for the sandwichlike trimethylenemethane complexes of group 8 (η6-C6H6)M-TMM (BzM-TMM), where M = Fe, Ru, Os, group 9 (η5-C5H5)M-TMM (CpM-TMM), where M = Co, Rh, Ir, and group 10, (η4-C4H4)M-TMM (CbM-TMM), where M = Ni, Pd, Pt. The nature of the metal–TMM bonding has been investigated with charge and energy decomposition analyses. The geometry optimization of the complexes gives sandwichlike structures where the terminal carbon atoms of the TMM ligands have significantly longer distances to the metal than to the central carbon. The calculated bond dissociations energies De of the TMM ligand are between 89.9 and 153.0 kcal/mol. The intrinsic interaction energies ΔEint and the De values for the heavier group 8 and group 9 elements become larger for the heavier atoms in the order BzFe-TMM < BzRu-TMM < BzOs-TMM and CpCoTMM < CpRh-TMM < CpIr-TMM, respectively. The group 10 elements exhibit a V-shaped trend for ΔEint and the De values with the sequence CbPd-TMM < CbNi-TMM < CbPt-TMM. The analysis of the bonding situation shows that the dominant orbital interactions come from the degenerate π interactions between the singly occupied degenerate π orbitals of the metal fragment and the TMM ligand. The degenerate π orbital of TMM has coefficients only at the terminal atom. In agreement with the shape of the orbitals, the EDA-NOCV method suggests that the metal–TMM bonding takes place mainly between the terminal carbon atoms and the metal while the σ bonding between the central carbon atom and the metal is rather weak. In contrast, the AIM analysis gives a bond path only between the metal atom and the closer central carbon atom but not to the more distant carbon atoms. This clearly shows that the AIM analysis does not faithfully represent the strongest pairwise interactions between the atoms in a molecule. The EDA-NOCV and the NBO methods agree that the TMM ligand in the complex carries only a small partial charge, which may be negative or positive.
Co-reporter:Terrance J. Hadlington;Markus Hermann;Jiaye Li;Dr. Gernot Frenking; Cameron Jones
Angewandte Chemie International Edition 2013 Volume 52( Issue 39) pp:10199-10203
Publication Date(Web):
DOI:10.1002/anie.201305689
Co-reporter:Dr. Kartik Chra Mondal; Herbert W. Roesky;Martin C. Schwarzer; Gernot Frenking;Benedikt Niepötter;Hilke Wolf;Dr. Regine Herbst-Irmer; Dietmar Stalke
Angewandte Chemie International Edition 2013 Volume 52( Issue 10) pp:2963-2967
Publication Date(Web):
DOI:10.1002/anie.201208307
Co-reporter:Nicole Holzmann;Deepak Dange;Dr. Cameron Jones;Dr. Gernot Frenking
Angewandte Chemie International Edition 2013 Volume 52( Issue 10) pp:3004-3008
Publication Date(Web):
DOI:10.1002/anie.201206305
Co-reporter:Dr. Kartik Chra Mondal;Dr. Herbert W. Roesky;Martin C. Schwarzer;Dr. Gernot Frenking;Dr. Igor Tkach;Hilke Wolf;Daniel Kratzert;Dr. Regine Herbst-Irmer;Benedikt Niepötter;Dr. Dietmar Stalke
Angewandte Chemie International Edition 2013 Volume 52( Issue 6) pp:
Publication Date(Web):
DOI:10.1002/anie.201210325
Co-reporter:Shiblee R. Barua; Wesley D. Allen; Elfi Kraka;Paul Jerabek;Rebecca Sure; Gernot Frenking
Chemistry - A European Journal 2013 Volume 19( Issue 47) pp:15941-15954
Publication Date(Web):
DOI:10.1002/chem.201302181
Abstract
The ground electronic state of C(BH)2 exhibits both a linear minimum and a peculiar angle-deformation isomer with a central B-C-B angle near 90°. Definitive computations on these species and the intervening transition state have been executed by means of coupled-cluster theory including single and double excitations (CCSD), perturbative triples (CCSD(T)), and full triples with perturbative quadruples (CCSDT(Q)), in concert with series of correlation-consistent basis sets (cc-pVXZ, X=D, T, Q, 5, 6; cc-pCVXZ, X=T, Q). Final energies were pinpointed by focal-point analyses (FPA) targeting the complete basis-set limit of CCSDT(Q) theory with auxiliary core correlation, relativistic, and non-Born–Oppenheimer corrections. Isomerization of the linear species to the bent form has a minuscule FPA reaction energy of 0.02 kcal mol−1 and a corresponding barrier of only 1.89 kcal mol−1. Quantum tunneling computations reveal interconversion of the two isomers on a timescale much less than 1 s even at 0 K. Highly accurate CCSD(T)/cc-pVTZ and composite c∼CCSDT(Q)/cc-pCVQZ anharmonic vibrational frequencies confirm matrix-isolation infrared bands previously assigned to linear C(BH)2 and provide excellent predictions for the heretofore unobserved bent isomer. Chemical bonding in the C(BH)2 species was exhaustively investigated by the atoms-in-molecules (AIM) approach, molecular orbital plots, various population analyses, local mode vibrations and force constants, unified reaction valley analysis (URVA), and other methods. Linear C(BH)2 is a cumulene, whereas bent C(BH)2 is best characterized as a carbene with little carbone character. Weak B–B attraction is clearly present in the unusual bent isomer, but its strength is insufficient to form a CB2 ring with a genuine boron–boron bond and attendant AIM bond path.
Co-reporter:Majid El-Hamdi, Miquel Solà, Gernot Frenking, and Jordi Poater
The Journal of Physical Chemistry A 2013 Volume 117(Issue 33) pp:8026-8034
Publication Date(Web):July 23, 2013
DOI:10.1021/jp4051403
A comparison between alkalimetal (M = Li, Na, K, and Rb) and group 11 transition metal (M = Cu, Ag, and Au) (MX)4 tetramers with X = H, F, Cl, Br, and I has been carried out by means of the Amsterdam Density Functional software using density functional theory at the BP86/QZ4P level of theory and including relativistic effects through the ZORA approximation. We have obtained that, in the case of alkalimetals, the cubic isomer of Td geometry is more stable than the ring structure with D4h symmetry, whereas in the case of group 11 transition metal tetramers, the isomer with D4h symmetry (or D2d symmetry) is more stable than the Td form. To better understand the results obtained we have made energy decomposition analyses of the tetramerization energies. The results show that in alkalimetal halide and hydride tetramers, the cubic geometry is the most stable because the larger Pauli repulsion energies are compensated by the attractive electrostatic and orbital interaction terms. In the case of group 11 transition metal tetramers, the D4h/D2d geometry is more stable than the Td one due to the reduction of electrostatic stabilization and the dominant effect of the Pauli repulsion.
Co-reporter:Dr. Mehmet Ali Celik;Dr. Gernot Frenking;Dr. Bernhard Neumüller;Dr. Wolfgang Petz
ChemPlusChem 2013 Volume 78( Issue 9) pp:1024-1032
Publication Date(Web):
DOI:10.1002/cplu.201300169
Abstract
Quantum chemical calculations at the BP86/TZVPP//BP86/SVP level of theory have been performed for the isoelectronic series of compounds [(PPh3)2CEH2]q (Eq=Be, B+, C2+, N3+, O4+). The equilibrium geometries and bond dissociation energies were calculated and the nature of the CE bond was investigated with charge and energy decomposition methods. The dication [(PPh3)2CCH2]2+ could become isolated as a salt compound with two counter ions [AlBr4]−. The X-ray structure analysis of [(PPh3)2CCH2]2+ gave bond lengths and angles that are in good agreement with the calculated data. The geometry optimization of [(PPh3)2COH2]4+ gave [(PPh3)2COH]3+ as the equilibrium structure. Bonding analysis of [(PPh3)2CEH2]q shows that [(PPh3)2CBeH2] and [(PPh3)2CBH2]+ possess donor–acceptor bonds in which the σ and π lone-pair electrons of (PPh3)2C donate into the vacant orbitals of the acceptor fragment. The multiply charged compounds are better described as substituted olefins [(PPh3)2CCH2]2+, [(PPh3)2CNH2]3+, and [(PPh3)2COH]3+, which possess electron-sharing σ and π bonds that arise from the interaction between the triplet states of [(PPh3)2C]2+ and the respective fragment CH2, (NH2)+, and (OH)+. The multiply charged cations [(PPh3)2CCH2]2+, [(PPh3)2CNH2]3+, and [(PPh3)2COH]3+ are calculated to be stable toward dissociation.
Co-reporter: Gernot Frenking;Markus Hermann
Angewandte Chemie 2013 Volume 125( Issue 23) pp:6036-6039
Publication Date(Web):
DOI:10.1002/ange.201301485
Co-reporter:Cameron Jones, Anastas Sidiropoulos, Nicole Holzmann, Gernot Frenking and Andreas Stasch
Chemical Communications 2012 vol. 48(Issue 79) pp:9855-9857
Publication Date(Web):16 Aug 2012
DOI:10.1039/C2CC35228A
Reduction of an N-heterocyclic carbene (NHC) adduct of SnCl2, viz. [(IPr)SnCl2] (IPr = :C{N(Dip)C(H)}2; Dip = 2,6-diisopropylphenyl), with a magnesium(I) dimer, has afforded the first NHC complex of a row 5 element in its diatomic form, [(IPr)SnSn(IPr)]; a computational analysis of the complex indicates that it comprises a singlet state, doubly bonded tin(0) fragment, :SnSn:, datively bonded by two NHC ligands.
Co-reporter:Rajendra S. Ghadwal, Ramachandran Azhakar, Herbert W. Roesky, Kevin Pröpper, Birger Dittrich, Catharina Goedecke and Gernot Frenking
Chemical Communications 2012 vol. 48(Issue 66) pp:8186-8188
Publication Date(Web):02 Jul 2012
DOI:10.1039/C2CC32887A
Formyl chloride (H(Cl)CO) is unstable at room temperature and decomposes to HCl and CO. Silicon analogue of formyl chloride, silaformyl chloride IPr·SiH(Cl)O·B(C6F5)3 (3) (IPr = 1,3-bis(2,6-diisopropyl-phenyl)imidazol-2-ylidene), was stabilized by Lewis donor–acceptor ligands. Compound 3 is not only the first stable acyclic silacarbonyl compound but also the first silacarbonyl halide reported so far.
Co-reporter:Catharina Goedecke, René Sitt, and Gernot Frenking
Inorganic Chemistry 2012 Volume 51(Issue 21) pp:11259-11265
Publication Date(Web):October 15, 2012
DOI:10.1021/ic301722q
Quantum chemical calculations using density functional theory at the BP86/TZ2P+ level and ab initio calculations at MP2/def2-TZVPP have been carried out for the donor–acceptor complexes [D→C6F4→BF3] (D = Xe, CO, N2,) and the dication [Xe→C6F4←Xe]2+. The calculations predict rather short D→C6F4(BF3) and (D)C6F4→BF3 bonds in the neutral systems which indicate rather strong binding interactions. The calculated partial charges which give large positive values for the donor moieties and negative values for the acceptor fragments and the large bond indices also suggest very strong donor–acceptor interactions D→C6F4→BF3 and Xe→C6F42+←Xe. An energy decomposition analysis suggests very strong intrinsic interactions for both systems. The donor–acceptor bonds in [D→C6F4→BF3] are much stronger than the direct donor–acceptor interactions D→BF3 which are only weakly bonded van der Waals complexes. The calculated donor–acceptor interactions D→C6F4(BF3) are 26.1 kcal/mol for D = Xe, 121.5 kcal/mol for D = CO, and 86.9 kcal/mol for D = N2. The strength of the intrinsic (D)C6F4→BF3 interactions are calculated to be between 51.1–51.6 kcal/mol. The theoretical bond dissociation energies for the decomposition of [D→C6F4→BF3] yielding D + C6F4 + BF3 suggests that the xenon compound [Xe→C6F4→BF3] is metastable but may become stabilized in the condensed phase by intermolecular interactions. The complexes [OC→C6F4→BF3] and [N2→C6F4→BF3] are predicted to be thermodynamically stable. It is suggested that the above adducts are examples of spacer-separated donor–acceptor complexes [D→S→A] which are a hitherto unrecognized class of molecules.
Co-reporter:Dr. Deepa Devarajan ;Dr. Gernot Frenking
Chemistry – An Asian Journal 2012 Volume 7( Issue 6) pp:1296-1311
Publication Date(Web):
DOI:10.1002/asia.201200022
Abstract
The singlet potential-energy surface (PES) of the system involving the atoms H, X, and E (the (H, X, E) system) in which X=N–Bi and E=C–Pb has been explored at the CCSD(T)/TZVPP and BP86/TZ2P+ levels of theory. The nature of the XE bonding has been analyzed with charge- and energy-partitioning methods. The calculations show that the linear isomers of the nitrogen systems lin-HEN and lin-HNE are minima on the singlet PES. The carbon compound lin-HCN (HCN=hydrogen cyanide) is 14.9 kcal mol−1 lower in energy than lin-HNC but the heavier group 14 homologues lin-HEN (E=Si–Pb) are between 64.8 and 71.5 kcal mol−1 less stable than the lin-HNE isomers. The phosphorous system (H, P, E) exhibits significant differences concerning the geometry and stability of the equilibrium structures compared with the nitrogen system. The linear form lin-HEP of the former system is much more stable than lin-HPE. The molecule lin-HCP is the only minimum on the singlet PES. It is 78.5 kcal mol−1 lower in energy than lin-HPC, which is a second-order saddle point. The heavier homologues lin-HPE, in which E=Si–Pb, are also second-order saddle points, whereas the bent-HPE structures are the global minima on the PES. They are between 10.3 (E=Si) and 36.5 kcal mol−1 (E=Pb) lower in energy than lin-HEP. The bent-HPE structures possess rather acute bending angles H-P-E between 60.1 (E=Si) and 79.7° (E=Pb). The energy differences between the heavier group 15 isomers lin-HEX (X=P–Bi) and the bent structures bent-HXE become continuously smaller. The silicon species lin-HSiBi is even 3.1 kcal mol−1 lower in energy than bent-HBiSi. The bending angle H-X-E becomes more acute when X becomes heavier. The drastic energy differences between the isomers of the system (H, X, E) are explained with three factors that determine the relative stabilities of the energy minima: 1) The different bond strength between the hydrogen bonds HX and HE. 2) The electronic excitation energy of the fragment HE from the X 2Π ground state to the 4Σ− excited state, which is required to establish a E≡X triple bond in the molecules lin-HEX. 3) The strength of the intrinsic XE interactions in the molecules. The trends of the geometries and relative energies of the linear, bent, and cyclic isomers are explained with an energy-decomposition analysis that provides deep insight into the nature of the bonding situation.
Co-reporter:Dr. Timo Bollermann;Mariusz Molon;Dr. Christian Gemel;Kerstin Freitag;Dr. Rüdiger W. Seidel;Dr. Moritz vonHopffgarten;Paul Jerabek;Dr. Gernot Frenking;Dr. Rol A. Fischer
Chemistry - A European Journal 2012 Volume 18( Issue 16) pp:4909-4915
Publication Date(Web):
DOI:10.1002/chem.201102758
Abstract
The synthesis, characterization, and theoretical investigation by means of quantum-chemical calculations of an oligonuclear metal-rich compound are presented. The reaction of homoleptic dinuclear palladium compound [Pd2(μ-GaCp*)3(GaCp*)2] with ZnMe2 resulted in the formation of unprecedented ternary Pd/Ga/Zn compound [Pd2Zn6Ga2(Cp*)5(CH3)3] (1), which was analyzed by 1H and 13C NMR spectroscopy, MS, elemental analysis, and single-crystal X-ray diffraction. Compound 1 consisted of two Cs-symmetric molecular isomers, as revealed by NMR spectroscopy, at which distinct site-preferences related to the Ga and Zn positions were observed by quantum-chemical calculations. Structural characterization of compound 1 showed significantly different coordination environments for both palladium centers. Whilst one Pd atom sat in the central of a bi-capped trigonal prism, thereby resulting in a formal 18-valence electron fragment, {Pd(ZnMe)2(ZnCp*)4(GaMe)}, the other Pd atom occupied one capping unit, thereby resulting in a highly unsaturated 12-valence electron fragment, {Pd(GaCp*)}. The bonding situation, as determined by atoms-in-molecules analysis (AIM), NBO partial charges, and molecular orbital (MO) analysis, pointed out that significant PdPd interactions had a large stake in the stabilization of this unusual molecule. The characterization and quantum-chemical calculations of compound 1 revealed distinct similarities to related M/Zn/Ga Hume–Rothery intermetallic solid-state compounds, such as Ga/Zn-exchange reactions, the site-preferences of the Zn/Ga positions, and direct MM bonding, which contributes to the overall stability of the metal-rich compound.
Co-reporter:Dr. Nozomi Takagi;Dr. Ralf Tonner ;Dr. Gernot Frenking
Chemistry - A European Journal 2012 Volume 18( Issue 6) pp:1772-1780
Publication Date(Web):
DOI:10.1002/chem.201100494
Abstract
Quantum-chemical calculations at the BP86/TZVPP level have been carried out for the heavy Group 14 homologues of carbodiphosphorane E(PPh3)2, where E=Si, Ge, Sn, Pb, which are experimentally unknown so far. The results of the theoretical investigation suggest that the tetrelediphosphoranes E(PPh3)2 (1 E) are stable compounds that could become isolated in a condensed phase. The molecules possess donor–acceptor bonds Ph3PEPPh3 to a bare tetrele atom E, which retains its four valence electrons as two electron lone pairs. The analysis of the bonding situation and the calculation of the chemical reactivity indicate that the molecules 1 E belong to the class of divalent E(0) compounds (ylidones). All molecules 1 C–1 Pb have very large first but also very large second proton affinities, which distinguishes them from the N-heterocyclic carbene homologues, in which the donor atom is a divalent E(II) species that possesses only one electron lone pair. Compounds 1 E are powerful double donors that strongly bind Lewis acids such as BH3 and AuCl in the complexes 1 E(BH3)n and 1 E(AuCl)n (n=1, 2). The bond dissociation energies (BDEs) of the second BH3 and AuCl molecules are only slightly less than the BDE of the first BH3 and AuCl. The results of this work are a challenge for experimentalists.
Co-reporter:Charity Flener Lovitt ; Gernot Frenking ;Gregory S. Girolami
Organometallics 2012 Volume 31(Issue 11) pp:4122-4132
Publication Date(Web):May 30, 2012
DOI:10.1021/om200456j
Density functional theory (DFT) is used to investigate the geometries and metal–ligand bonding in nickel complexes of bidentate phosphines, NiX2(R2P(CH2)nPR2), where X = H, CO, n = 1–3, and R = H, Me, CF3, Et, i-Pr, t-Bu, Ph, OMe, F. The net donor–acceptor properties of the phosphine ligands can be deduced from the computed frequency of the symmetric CO stretch of the Ni(CO)2(R2P(CH2)nPR2) carbonyl complexes. This frequency (in cm–1) can be estimated from the empirical expression ν(CO) = 1988 + ∑χB – 4n, where the sum is over the four substituents on the bidentate phosphine, χB is a substituent-dependent parameter, and n is the number of carbon atoms in the backbone (1 ≤ n ≤ 3). The deduced values of χB (in units of cm–1)—t-Bu (0.0), i-Pr (0.8), Et (3.0), Me (4.0), Ph (4.3), H (6.3), OMe (10.8), CF3 (17.8), and F (18.3)—are generally similar to Tolman’s electronic parameter χ derived from nickel complexes of unidentate phosphines. For the NiH2(R2P(CH2)nPR2) hydride complexes, the global minimum is a nonclassical dihydrogen structure, irrespective of the nature of the phosphine. For bidentate phosphines that are strongly donating, a classical cis-dihydride structure lies higher in energy (in some cases, by only 0.4 kcal mol–1 above the global minimum). For phosphines that are less electron donating, the dihydride structure is no longer a local minimum but instead is an inflection point on the potential energy surface. Atoms in molecules (AIM) and natural bond order (NBO) analyses confirm that the nickel–dihydrogen interaction involves a three-center–two-electron bond. The Kohn–Sham molecular orbital diagram and energy decomposition analysis of these complexes show that metal to H2 π back-donation is the dominant orbital component for phosphines with electron-donating substituents, whereas H2 to metal σ donation is dominant for phosphines with electron-withdrawing substituents. The EDA results clearly indicate that long H–H distances are seen when the metal to H2 π back-donation dominates over H2 to M σ donation.
Co-reporter:Thi Ai Nhung Nguyen ;Dr. Gernot Frenking
Chemistry - A European Journal 2012 Volume 18( Issue 40) pp:12733-12748
Publication Date(Web):
DOI:10.1002/chem.201200741
Abstract
Quantum chemical calculations at the BP86/TZVPP//BP86/SVP level are performed for the tetrylone complexes [W(CO)5-E(PPh3)2] (W-1 E) and the tetrylene complexes [W(CO)5-NHE] (W-2 E) with E=C–Pb. The bonding is analyzed using charge and energy decomposition methods. The carbone ligand C(PPh3) is bonded head-on to the metal in W-1 C, but the tetrylone ligands E(PPh3)2 are bonded side-on in the heavier homologues W-1 Si to W-1 Pb. The WE bond dissociation energies (BDEs) increase from the lighter to the heavier homologues (W-1 C: De=25.1 kcal mol−1; W-1 Pb: De=44.6 kcal mol−1). The W(CO)5C(PPh3)2 donation in W-1 C comes from the σ lone-pair orbital of C(PPh3)2, whereas the W(CO)5E(PPh3)2 donation in the side-on bonded complexes with E=Si–Pb arises from the π lone-pair orbital of E(PPh3)2 (the HOMO of the free ligand). The π-HOMO energy level rises continuously for the heavier homologues, and the hybridization has greater p character, making the heavier tetrylones stronger donors than the lighter systems, because tetrylones have two lone-pair orbitals available for donation. Energy decomposition analysis (EDA) in conjunction with natural orbital for chemical valence (NOCV) suggests that the WE BDE trend in W-1 E comes from the increase in W(CO)5E(PPh3)2 donation and from stronger electrostatic attraction, and that the E(PPh3)2 ligands are strong σ-donors and weak π-donors. The NHE ligands in the W-2 E complexes are bonded end-on for E=C, Si, and Ge, but side-on for E=Sn and Pb. The WE BDE trend is opposite to that of the W-1 E complexes. The NHE ligands are strong σ-donors and weak π-acceptors. The observed trend arises because the hybridization of the donor orbital at atom E in W-2 E has much greater s character than that in W-1 E, and even increases for heavier atoms, because the tetrylenes have only one lone-pair orbital available for donation. In addition, the WE bonds of the heavier systems W-2 E are strongly polarized toward atom E, so the electrostatic attraction with the tungsten atom is weak. The BDEs calculated for the WE bonds in W-1 E, W-2 E and the less bulky tetrylone complexes [W(CO)5-E(PH3)2] (W-3 E) show that the effect of bulky ligands may obscure the intrinsic WE bond strength.
Co-reporter:Dr. Mehmet Ali Celik;Rebecca Sure;Dr. Susanne Klein;Dr. Rei Kinjo;Dr. Guy Bertr;Dr. Gernot Frenking
Chemistry - A European Journal 2012 Volume 18( Issue 18) pp:5676-5692
Publication Date(Web):
DOI:10.1002/chem.201103965
Abstract
Quantum chemical calculations using DFT (BP86, M05-2X) and ab initio methods (CCSD(T), SCS-MP2) have been carried out on the borylene complexes (BH)L2 and nitrogen cation complexes (N+)L2 with the ligands L=CO, N2, PPh3, NHCMe, CAAC, and CAACmodel. The results are compared with those obtained for the isoelectronic carbones CL2. The geometries and bond dissociation energies of the ligands, the proton affinities, and adducts with the Lewis acids BH3 and AuCl were calculated. The nature of the bonding has been analyzed with charge and energy partitioning methods. The calculated borylene complexes (BH)L2 have trigonal planar coordinated boron atoms which possess rather short BL bonds. The calculated bond dissociation energies (BDEs) of the ligands for complexes where L is a carbene (NHC or CAAC) are very large (De=141.6–177.3 kcal mol−1) which suggest that such species might become isolated in a condensed phase. The borylene complexes (BH)(PPh3)2 and (BH)(CO)2 have intermediate bond strengths (De=90.1 and 92.6 kcal mol−1). Substituted homologues with bulky groups at boron which protect the boron atom from electrophilic attack might also be stable enough to become isolated. The BDE of (BH)(N2)2 is much smaller (De=31.9 kcal mol−1), but could become observable in a low-temperature matrix. The proton affinities of the borylene complexes are very large, particularly for the bulky adducts with L=PPh3, NHCMe, CAACmodel and CAAC and thus, they are superbases. All (BH)L2 molecules bind strongly AuCl either η1 (L=N2, PPh3, NHCMe, CAAC) or η2 (L=CO, CAACmodel). The BDEs of H3B(BH)L2 adducts which possess a hitherto unknown boronboron donor–acceptor bond are smaller than for the AuCl complexes. The strongest bonded BH3 adduct that might be isolable is (BH)(PPh3)2BH3 (De=36.2 kcal mol−1). The analysis of the bonding situation reveals that (BH)L2 bonding comes mainly from the orbital interactions which has three major contributions, that is, the donation from the symmetric (σ) and antisymmetric (π||) combination of the ligand lone-pair orbitals into the vacant MOs of BH L(BH)L and the L(BH)L π backdonation from the boron lone-pair orbital. The nitrogen cation complexes (N+)L2 have strongly bent LNL geometries, in which the calculated bending angle varies between 113.9° (L=N2) and 146.9° (L=CAAC). The BDEs for (N+)L2 are much larger than those of the borylene complexes. The carbene ligands NHC and CAAC but also the phosphane ligands PPh3 bind very strongly between De=358.4 kcal mol−1 (L=PPh3) and De=412.5 kcal mol−1 (L=CAACmodel). The proton affinities (PA) of (N+)L2 are much smaller and they bind AuCl and BH3 less strongly compared with (BH)L2. However, the PAs (N+)L2 for complexes with bulky ligands L are still between 139.9 kcal mol−1 (L=CAACmodel) and 168.5 kcal mol−1 (L=CAAC). The analysis of the (N+)L2 bonding situation reveals that the binding interactions come mainly from the L(N+)L donation while L(N+)L π backdonation is rather weak.
Co-reporter:Jiaye Li;Markus Hermann;Dr. Gernot Frenking; Cameron Jones
Angewandte Chemie International Edition 2012 Volume 51( Issue 34) pp:8611-8614
Publication Date(Web):
DOI:10.1002/anie.201203607
Co-reporter:Jiaye Li;Markus Hermann;Dr. Gernot Frenking; Cameron Jones
Angewandte Chemie 2012 Volume 124( Issue 34) pp:8739-8742
Publication Date(Web):
DOI:10.1002/ange.201203607
Co-reporter:Catharina Goedecke ; Michael Leibold ; Ulrich Siemeling
Journal of the American Chemical Society 2011 Volume 133(Issue 10) pp:3557-3569
Publication Date(Web):February 18, 2011
DOI:10.1021/ja109812r
Quantum-chemical calculations using DFT and ab initio methods have been carried out for 32 carbenes RR′C which comprise different classes of compounds and the associated ketenes RR′C═C═O. The calculated singlet−triplet gaps ΔES−T of the carbenes exhibit a very high correlation with the bond dissociation energies (BDEs) of the ketenes. An energy decomposition analysis of the RR′C−CO bond using the triplet states of the carbene and CO as interacting fragments supports the assignment of ΔES−T as the dominant factor for the BDE but also shows that the specific interactions of the carbene may sometimes compensate for the S/T gap. The trend of the interaction energy ΔEint values is mainly determined by the Pauli repulsion between the carbene and CO. The stability of amino-substituted ketenes strongly depends on the destabilizing conjugation between the nitrogen lone-pair orbital and the ketene double bonds. There is a ketene structure of the unsaturated N-heterocyclic carbene parent compound NHC1 with CO as a local energy minimum on the potential-energy surface. However, the compound NHC1−CO is thermodynamically unstable toward dissociation. The saturated homologue NHC2−CO has only a very small bond dissociation energy of De = 3.2 kcal/mol. The [3]ferrocenophane-type compound FeNHC−CO has a BDE of De = 16.0 kcal/mol.
Co-reporter:Rajendra S. Ghadwal ; Ramachandran Azhakar ; Herbert W. Roesky ; Kevin Pröpper ; Birger Dittrich ; Susanne Klein
Journal of the American Chemical Society 2011 Volume 133(Issue 44) pp:17552-17555
Publication Date(Web):October 14, 2011
DOI:10.1021/ja206702e
A stable silicon analogue of an acid anhydride {PhC(ButN)2}Si{═O·B(C6F5)3}O–Si(H){═O·B(C6F5)3}{(NBut)(HNBut)CPh} (4) with a O═Si–O–Si═O core has been prepared by treating monochlorosilylene PhC(ButN)2SiCl (1) with H2O·B(C6F5)3 in the presence of NHC (NHC = 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene). Compound 4 has been characterized by elemental analysis and multinuclear NMR spectroscopic investigations. The molecular structure of 4 has been established by single-crystal X-ray diffraction studies, and DFT calculations support the experimental results.
Co-reporter:Jiaye Li ; Christian Schenk ; Catharina Goedecke ; Gernot Frenking ;Cameron Jones
Journal of the American Chemical Society 2011 Volume 133(Issue 46) pp:18622-18625
Publication Date(Web):October 25, 2011
DOI:10.1021/ja209215a
The reduction of the bulky amido-germanium(II) chloride complex, LGeCl (L = N(SiMe3)(Ar*); Ar* = C6H2Me{C(H)Ph2}2-4,2,6), with the magnesium(I) dimer, [{(MesNacnac)Mg}2] (MesNacnac = [(MesNCMe)2CH]−; Mes = mesityl), afforded LGeGeL, which represents the first example of a digermyne with a Ge–Ge single bond. Computational studies of the compound have highlighted significant electronic differences between it and multiply bonded digermynes. LGeGeL was shown to cleanly activate H2 in solution or the solid state, at temperatures as low as −10 °C, to give the mixed valence compound, LGeGe(H)2L.
Co-reporter:Cameron Jones, Simon J. Bonyhady, Nicole Holzmann, Gernot Frenking, and Andreas Stasch
Inorganic Chemistry 2011 Volume 50(Issue 24) pp:12315-12325
Publication Date(Web):May 6, 2011
DOI:10.1021/ic200682p
A synthetic route to the new amidine (DipNH)(DipN)C(C6H4But-4) (ButisoH; Dip = C6H3Pri2-2,6) has been developed. Its deprotonation with either LiBun or KN(SiMe3)2 yields the amidinate complexes [M(Butiso)] (M = Li or K). Their reactions with group 14 element halides/pseudohalides afford the heteroleptic group 14 complexes [(Butiso)SiCl3], [(Butiso)ECl] (E = Ge or Sn), and [{(Butiso)Pb(μ-O3SCF3)(THF)}∞], all of which have been crystallographically characterized. In addition, the synthesis and spectroscopic characterization of the homoleptic complex [Pb(Butiso)2] is reported. Reductions of the heteroleptic complexes with a soluble magnesium(I) dimer, [{(MesNacnac)Mg}2] (MesNacnac = [(MesNCMe)2CH]−; Mes = mesityl), have given moderate-to-high yields of the group 14 element(I) dimers [{(Butiso)E}2] (E = Si, Ge, or Sn), the X-ray crystallographic studies of which reveal trans-bent structures. The corresponding lead(I) complex could not be prepared. Comprehensive spectroscopic and theoretical analyses of [{(Butiso)E}2] have allowed their properties to be compared. All complexes possess E–E single bonds and can be considered as intramolecularly base-stabilized examples of ditetrelynes, REER. Taken as a whole, this study highlights the synthetic utility of soluble and easy to prepare magnesium(I) dimers as valuable alternatives to the harsh, and often insoluble, alkali-metal reducing agents that are currently widely employed in the synthesis of low-oxidation-state organometallic/inorganic complexes.
Co-reporter:H. V. Rasika Dias ; Chandrakanta Dash ; Muhammed Yousufuddin ; Mehmet Ali Celik
Inorganic Chemistry 2011 Volume 50(Issue 10) pp:4253-4255
Publication Date(Web):April 20, 2011
DOI:10.1021/ic200757j
A cationic gold carbonyl complex has been synthesized and characterized using several techniques including X-ray crystallography. [(Mes3P)Au(CO)][SbF6] (Mes = 2,4,6-Me3C6H2) has a linear, two-coordinate gold atom. This compound displays the CO stretching frequency at 2185 cm–1. The 13C NMR signal of the gold-bound 13CO appears as a doublet centered at δ 182.6 (2JC,P = 115 Hz). A computational study shows that the Au–CO bond consists of electrostatic attraction, Au ← CO donation, and significant Au → CO π-back-bonding components. Polarization of the CO bond caused by the electrostatic effect of the cationic gold center is mainly responsible for the large blue shift in the CO stretching frequency.
Co-reporter:Nozomi Takagi ; Andreas Krapp
Inorganic Chemistry 2011 Volume 50(Issue 3) pp:819-826
Publication Date(Web):January 6, 2011
DOI:10.1021/ic101227u
The bonding situation of homonuclear and heteronuclear metal−metal multiple bonds in R3M−M′R3 (M, M′ = Cr, Mo, W; R = Cl, NMe2) is investigated by density functional theory (DFT) calculations, with the help of energy decomposition analysis (EDA). The M−M′ bond strength increases as M and M′ become heavier. The strongest bond is predicted for the 5d−5d tungsten complexes (NMe2)3W−W(NMe2)3 (De = 103.6 kcal/mol) and Cl3W−WCl3 (De = 99.8 kcal/mol). Although the heteronuclear molecules with polar M−M′ bonds are not known experimentally, the predicted bond dissociation energies of up to 94.1 kcal/mol for (NMe2)3Mo−W(NMe2)3 indicate that they are stable enough to be isolated in the condensed phase. The results of the EDA show that the stronger R3M−M′R3 bonds for heavier metal atoms can be ascribed to the larger electrostatic interaction caused by effective attraction between the expanding valence orbitals in one metal atom and the more positively charged nucleus in the other metal atom. The orbital interaction reveal that the covalency of the homonuclear and heteronuclear R3M−M′R3 bonds is due to genuine triple bonds with one σ- and one degenerate π-symmetric component. The metal−metal bonds may be classified as triple bonds where π-bonding is much stronger than σ-bonding; however, the largest attraction comes from the quasiclassical contribution to the metal−metal bonding. The heterodimetallic species show only moderate polarity and their properties and stabilities are intermediate between the corresponding homodimetallic species, a fact which should allow for the experimental isolation of heterodinuclear species. CASPT2 calculations of Cl3M−MCl3 (M = Cr, Mo, W) support the assignment of the molecules as triply bonded systems.
Co-reporter:Rajendra K. Jangid, Neelima Gupta, Raj K. Bansal, Moritz von Hopffgarten, Gernot Frenking
Tetrahedron Letters 2011 Volume 52(Issue 14) pp:1721-1724
Publication Date(Web):6 April 2011
DOI:10.1016/j.tetlet.2011.02.007
Theoretical calculations at the DFT (B3LYP/6-31+G**) level of the model Diels–Alder (DA) reactions of 1-methyl-3-(methoxycarbonyl)-2-phosphaindolizine with 1,3-butadiene in the presence of methylaluminium dichloride reveal that the co-ordination of organoaluminium reagent to the carbonyl oxygen increases the activation barrier compared to that for the uncomplexed 2-phosphaindolizine. On the other hand, co-ordination of the organoaluminium reagent to the σ2, λ3 P atom lowers the activation barrier by ∼6 kcal mol−1. 1-Methyl-2-phosphaindolizines having an electron-withdrawing group at the 3-position only undergo DA reaction with 2,3-dimethylbutadiene in the presence of the ethylaluminium dichloride catalyst in methylene chloride at a low temperature to afford [2+4] cycloadducts. The formation of an intermediate having the ethylaluminium reagent co-ordinated to σ2, λ3 P atom has been detected by 31P NMR. The products have been characterized by 31P and 1H NMR studies.
Co-reporter:Neelima Gupta;Raj K. Bansal;Moritz von Hopffgarten
Journal of Physical Organic Chemistry 2011 Volume 24( Issue 9) pp:786-797
Publication Date(Web):
DOI:10.1002/poc.1829
The competitive 1,5-electrocyclization versus intramolecular 1,5-proton shift in imidazolium allylides and imidazolium 2-phosphaallylides has been investigated theoretically at the DFT (B3LYP/6-311 + +G**//B3LYP/6-31G**) level. 1,5-Electrocyclization follows pericyclic mechanism and its activation barrier is lower than that for the pseudopericyclic mechanism by ∼5–6 kcal mol−1. The activation barriers for 1,5-electrocyclization of imidazolium 2-phosphaallylides are found to be smaller than those for their nonphosphorus analogues by ∼3–5 kcal mol−1. There appears to be a good correlation between the activation barrier for intramolecular 1,5-proton shift and the density of the negative charge at C8, except for the ylides having fluorine substituent at this position (7b and 8b). The presence of fluorine atom reduces the density of the negative charge at C8 (in 7b it becomes positively charged) and thus raises the activation barrier. The ylides 7f and 8f having CF3 group at C8, in preference to the 1,5-proton shift, follow an alternative route leading to different carbenes which is accompanied by the loss of HF. The carbenes Pr7,8b–e resulting from intramolecular 1,5-proton shift have a strong tendency to undergo intramolecular SN2 type reaction, the activation barrier being 7–28 kcal mol−1. Copyright © 2011 John Wiley & Sons, Ltd.
Co-reporter:Dipl.-Chem. Nicole Holzmann;Dr. Andreas Stasch; Cameron Jones; Gernot Frenking
Chemistry - A European Journal 2011 Volume 17( Issue 48) pp:13517-13525
Publication Date(Web):
DOI:10.1002/chem.201101915
Abstract
Quantum chemical calculations using density functional theory at the BP86/TZVPP level and ab initio calculations at the SCS-MP2/TZVPP level have been carried out for the group 13 complexes [(NHC)(EX3)] and [(NHC)2(E2Xn)] (E=B to In; X=H, Cl; n=4, 2, 0; NHC=N-heterocyclic carbene). The monodentate Lewis acids EX3 and the bidentate Lewis acids E2Xn bind N-heterocyclic carbenes rather strongly in donor–acceptor complexes [(NHC)(EX3)] and [(NHC)2(E2Xn)]. The equilibrium structures of the bidentate complexes depend on the electronic reference state of E2Xn, which may vary for different atoms E and X. All complexes [(NHC)2(E2X4)] possess Cs symmetry in which the NHC ligands bind in a trans conformation to the group 13 atoms E. The complexes [(NHC)2(E2H2)] with E=B, Al, Ga have also Cs symmetry with a trans arrangement of the NHC ligands and a planar CE(H)E(H)C moiety that has a EE π bond. In contrast, the indium complex [(NHC)2(In2H2)] has Ci symmetry with pyramidal-coordinated In atoms in which the hydrogen atoms are twisted above and below the CInInC plane. The latter Ci form is calculated for all chloride systems [(NHC)2(E2Cl2)], but the boron complex [(NHC)2(B2Cl2)] deviates only slightly from Cs symmetry. The B2 fragment in the linear coordinated complex [(NHC)2(B2)] has a highly excited (3)1Σg− reference state, which gives an effective B≡B triple bond with a very short interatomic distance. The heavier homologues [(NHC)2(E2)] (E=Al to In) exhibit a anti-periplanar arrangement of the NHC ligands in which the E2 fragments have a (1)1Δg reference state and an EE double bond. The calculated energies suggest that the dihydrogen release from the complexes [(NHC)(EH3)] and [(NHC)2(E2Hn)] becomes energetically more favourable when atom E becomes heavier. The indium complexes should therefore be the best candidates of the investigated series for hydrogen-storage systems that could potentially deliver dihydrogen at close to ambient temperature. The hydrogenation reaction of the dimeric magnesium(I) compound [LMgMgL] (L=β-diketiminate) with [(NHC)(EH3)] becomes increasingly exothermic with the trend B<Al<Ga<In.
Co-reporter:Moritz von Hopffgarten and Gernot Frenking
The Journal of Physical Chemistry A 2011 Volume 115(Issue 45) pp:12758-12768
Publication Date(Web):August 9, 2011
DOI:10.1021/jp2038762
The bonding situation of the icosahedral compounds [M(EH)12] (M = Cr, Mo, W; E = Zn, Cd, Hg), which are model systems for the isolated species [Mo(ZnCp*)3(ZnMe)9] possessing the coordination number 12 at the central atom M, have been analyzed with a variety of charge and energy decomposition methods (AIM, EDA-NOCV, WBI, MO). The results give a coherent picture of the electronic structure and the nature of the interatomic interactions. The compounds [M(EH)12] are transition metal complexes that possess 12 M-EH radial bond paths (AIM) that can be described as 6 three-center two-electron bonds (MO). The radial M-EH bonds come from the electron sharing interactions mainly between the singly occupied valence s and d AOs of the central atom M and the singly occupied EH valence orbitals (MO, EDA-NOCV). The orbital interactions provide ∼42% of the total attraction, while the electrostatic attraction contributes ∼58% to the metal–ligand bonding (EDA-NOCV). There is a weak peripheral E–E bonding in [M(EH)12] that explains the unusually high coordination number (MO). The peripheral bonding leads for some compounds [M(EH)12] to the emergence of E–E bond paths, while in others it does not (AIM). The relative strength of the radial and peripheral bonding in [Al13]− and [Pt@Pb12]2– is clearly different from the situation in [M(EH)12], which supports the assignments of the former species as cluster compounds or inclusion compounds (MO, WBI). The bonding situation in [WAu12] is similar to that in [M(EH)12].
Co-reporter:Timo Bollermann;Kerstin Freitag;Dr. Christian Gemel;Dr. Rüdiger W. Seidel;Moritz vonHopffgarten;Dr. Gernot Frenking;Dr. Rol A. Fischer
Angewandte Chemie 2011 Volume 123( Issue 3) pp:798-802
Publication Date(Web):
DOI:10.1002/ange.201005808
Co-reporter:Taka Shimizu
Theoretical Chemistry Accounts 2011 Volume 130( Issue 2-3) pp:269-277
Publication Date(Web):2011 October
DOI:10.1007/s00214-011-0974-0
Quantum chemical calculations using DFT (BP86) and ab initio methods (MP2, MP4 and CCSD(T)) have been carried out for the title compounds. The nature of the Pb–Pb interactions has been investigated with an energy decomposition analysis. The energy minimum structures of the halogen substituted Pb2X2 molecules possess a doubly bridged butterfly geometry A like the parent system Pb2H2. The unusual geometry can be explained with the interactions between PbX fragments in the X2Π ground state which leads to one Pb–Pb electron-sharing σ bond and two donor–acceptor bonds between the Pb–X bonds as donor and vacant p(π) AOs of Pb. The energy difference between the equilibrium form A and the linear structure XPb≡PbX (E) which is a second-order saddle point is much higher when X is a halogen atom than for X = H. This is because the a4Σ− ← X2Π excitation energies of PbX (X = F–I) are higher than for PbH. The structural isomers B, D1, D2, E, F1, F2 and G of Pb2X2 are no minima on the potential energy surface.
Co-reporter:Dr. Catharine Esterhuysen; Gernot Frenking
Chemistry - A European Journal 2011 Volume 17( Issue 36) pp:9944-9956
Publication Date(Web):
DOI:10.1002/chem.201101213
Abstract
Quantum chemical calculations have been performed for the dicoordinated carbon compounds C(PPh3)2, C(NHCMe)2, R2CC=CR2 (R=H, F, NMe2), C3O2, C(CN)2− and N-methyl-substituted N-heterocyclic carbene (NHCMe). The geometries of the complexes in which the dicoordinated carbon molecules bind as ligands to one and two AuCl moieties have been optimized and the strength and nature of the metal–ligand interactions in the mono- and diaurated complexes were investigated by means of energy decomposition analysis. The goal of the study is to elucidate the differences in the chemical behavior between carbones, allenes and carbenes. The results show that carbones bind one and two AuCl species in η1 fashion, whereas allenes bind them in η2 fashion. Compounds with latent divalent carbon(0) character can coordinate in more than one way, with the dominant mode indicating the degree of carbone or allene character. The calculated structures of the mono- and diaurated tetraaminoallenes (TAAs) reveal that TAAs exhibit a chameleon-like behavior: The bonding situation in the equilibrium structure is best described as allene [(R2N)2]CCC[(NR2)2] in which the central carbon atom is a tetravalent CIV species, but the reactivity suggests that TAAs should be considered as divalent C0 compounds C{C[(NR2)2]}2, that is, as “hidden” carbones. Carbon suboxide binds one AuCl preferentially in the η1 mode, whereas the equilibrium structures of the η1- and η2-bonded diaurated complex are energetically nearly degenerate. The doubly negatively charged isoelectronic carbone C(CN)22− binds one and two AuCl very strongly in characteristic η1 fashion. The N-heterocyclic carbene complex, [NHCMe(AuCl)], possesses a high bond dissociation energy (BDE) for the splitting off of AuCl. The diaurated NHC adduct, [NHCMe(AuCl)2], has two η1-bonded AuCl moieties that exhibit aurophilic attraction, which yield a moderate bond strength that might be large enough for synthesizing the complex. The BDE for the second AuCl in [NHCMe(AuCl)2] is clearly smaller than the values for the second AuCl in doubly aurated carbone complexes.
Co-reporter:Timo Bollermann;Kerstin Freitag;Dr. Christian Gemel;Dr. Rüdiger W. Seidel;Moritz vonHopffgarten;Dr. Gernot Frenking;Dr. Rol A. Fischer
Angewandte Chemie International Edition 2011 Volume 50( Issue 3) pp:772-776
Publication Date(Web):
DOI:10.1002/anie.201005808
Co-reporter:Dr. Rajendra S. Ghadwal;Dr. Herbert W. Roesky;Kevin Pröpper;Dr. Birger Dittrich;Susanne Klein;Dr. Gernot Frenking
Angewandte Chemie International Edition 2011 Volume 50( Issue 23) pp:5374-5378
Publication Date(Web):
DOI:10.1002/anie.201101320
Co-reporter:Ulrich Siemeling, Christian Färber, Clemens Bruhn, Michael Leibold, Detlef Selent, Wolfgang Baumann, Moritz von Hopffgarten, Catharina Goedecke and Gernot Frenking
Chemical Science 2010 vol. 1(Issue 6) pp:697-704
Publication Date(Web):22 Oct 2010
DOI:10.1039/C0SC00451K
N-Heterocyclic carbenes (NHCs) are extremely valuable as nucleophilic organocatalysts. They are widely applied as ligands in transition-metal catalysed reactions, where they are known as particularly potent σ-donors. They are commonly viewed as workhorses exhibiting reliable, but undramatic, chemical behaviour. The N → Ccarbene π-donation stabilises NHCs at the expense of low reactivity towards nucleophiles. In contrast to NHCs, stable (alkyl)(amino)carbenes exhibit spectacular reactivity, allowing, for example, the splitting of hydrogen and ammonia and the fixation of carbon monoxide. NHCs have been judged to be electronically not suitable for showing similar reactivity. Here, we demonstrate that a ferrocene-based NHC is able to add ammonia, methyl acrylate, tert-butyl isocyanide, and carbon monoxide—reactions typical of (alkyl)(amino)carbenes, but unprecedented for diaminocarbenes. We also show that even the simplest stable diaminocarbene, C(NiPr2)2, adds CO. This reaction affords a β-lactam by a subsequent intramolecular process involving a C–H activation. Our results shed new light on the chemistry of diaminocarbenes and offer great potential for synthetic chemistry and catalysis.
Co-reporter:Markus Rullich, Ralf Tonner and Gernot Frenking
New Journal of Chemistry 2010 vol. 34(Issue 8) pp:1760-1773
Publication Date(Web):05 Jul 2010
DOI:10.1039/C0NJ00208A
Quantum chemical calculations at the DFT (B3LYP) and ab initio level (CCSD(T)) have been carried out for the transition states and reaction products of the addition reactions of H2, NH3, CH4, H2O, C6H6, C2H6, C2H4, C2H2, CH3Cl, CH3F and SiH4 to model N-heterocyclic carbenes (NHCs) and P-heterocyclic carbenes (PHCs). The calculations show that PHCs have substantially lower activation barriers than NHCs for breaking the single bonds H–H, O–H, N–H, C–H, C–F, C–Cl and Si–H, as well as the π-bonds in benzene, ethylene and acetylene. The main reason for the higher reactivity of PHCs is their energetically lower-lying LUMO compared to NHCs. The energy level of the LUMO and the electrophilicity of PHCs strongly depends on pyramidalization at the carbene centre. Bulky ligands stabilize intrinsically unstable PHCs because they enforce a more planar arrangement at the carbene centre, which enhances the π-donation from the phosphorus lone-pair MO to the formally empty p(π) orbital at the divalent carbon atom. This raises the energy level of the LUMO but the higher reactivity of the PHC is preserved.
Co-reporter:Cherumuttathu H. Suresh and Gernot Frenking
Organometallics 2010 Volume 29(Issue 21) pp:4766-4769
Publication Date(Web):June 24, 2010
DOI:10.1021/om100260p
A comprehensive analysis of the nature of metal−carbon bonding in group 6 metallacyclobutadienes has been carried out by means of energy decomposition analysis and topological properties of electron density. The results suggest strong σ- and π-bonding interactions between the metal and the β-carbon (direct 1,3-metal−carbon bonding) as well as the existence of the β-carbon in a new type of planar tetracoordinate state characterized by a “catastrophe” bonding interaction with the metal. The importance of this bonding for alkyne metathesis is indicated.
Co-reporter:Dipl.-Chem. Susanne Klein ;Dr. Gernot Frenking
Angewandte Chemie 2010 Volume 122( Issue 39) pp:7260-7264
Publication Date(Web):
DOI:10.1002/ange.201002773
Co-reporter:Dipl.-Chem. Susanne Klein ;Dr. Gernot Frenking
Angewandte Chemie International Edition 2010 Volume 49( Issue 39) pp:7106-7110
Publication Date(Web):
DOI:10.1002/anie.201002773
Co-reporter:Timo Bollermann;Dr. Thomas Cadenbach;Dr. Christian Gemel;Moritz vonHopffgarten;Dr. Gernot Frenking;Dr. Rol A. Fischer
Chemistry - A European Journal 2010 Volume 16( Issue 45) pp:13372-13384
Publication Date(Web):
DOI:10.1002/chem.201001650
Abstract
The synthesis and structural characterization of novel, metal-rich, highly coordinated compounds [Mo(M′R)12] and [M(M′R)8] (M: Pd, Pt, Mo; M′: Zn, Cd; R: Me=CH3, Cp*=pentamethylcyclopentadienyl) are reported. Additionally, a description of the bonding situation of the new compounds by means of quantum-chemical calculations is presented including the Hg analogues. Reaction of [Pt(GaCp*)4] with CdMe2 results in the formation of the unprecedented all-Cd coordinated [Pt(CdMe)4(CdCp*)4] (1). Similarly, the treatment of the all-Zn coordinated [Pd(ZnMe)4(ZnCp*)4] with CdMe2 affords the novel Zn/Cd mixed compound [Pd(CdMe)4(ZnCp*)4] (2). The related Zn/Cd mixed compound [Mo(ZnCp*)3(CdMe)9] (3) is prepared by reaction of [Mo(ZnCp*)4(GaMe)4] with an excess amount of CdMe2. All compounds were analyzed by 1H and 13C NMR spectroscopy, elemental analysis, and single-crystal X-ray diffraction. The bonding situation of these highly coordinated, metal-rich molecules 1–3 were studied by quantum-chemical calculations using density functional theory (DFT) at the BP86/TZ2P+ level, atoms-in-molecules (AIM) analysis, and energy-decomposition analysis (EDA), as well as the its natural orbitals for chemical valence variation (EDA-NOCV) and including the hypothetically all-Hg-coordinated analogues. The results point out that the radial interactions M–M′ in the icosahedral compounds that have twelve ligands are best described as classical electron-pair-sharing covalent bonds, whereas the dodecahedral species, which have eight ligands, exhibit metal–ligand donor–acceptor bonds. The attractive interactions between the metal–ligand fragments M′R by means of M′–M′ bonds are weaker but not insignificant. All complexes fulfill the 18-electron rule. The analysis clarifies the electronic structures as being distinctly different from typical endohedral clusters M@(M′′R)n that exhibit strong peripheral M′′–M′′ interactions: The M′M′ bonds are not strong enough to yield stable (M′R)n cages.
Co-reporter:Susanne Klein Dipl.-Chem.;Ralf Tonner Dr.
Chemistry - A European Journal 2010 Volume 16( Issue 33) pp:10160-10170
Publication Date(Web):
DOI:10.1002/chem.201000174
Abstract
Quantum-chemical calculations using DFT and ab initio methods have been carried out for fourteen divalent carbon(0) compounds (carbones), in which the bonding situation at the two-coordinate carbon atom can be described in terms of donor–acceptor interactions LCL. The charge- and energy-decomposition analysis of the electronic structure of compounds 1–10 reveals divalent carbon(0) character in different degrees for all molecules. Carbone-type bonding LCL is particularly strong for the carbodicarbenes 1 and 2, for the “bent allenes” 3 a, 3 b, 4 a, and 4 b, and for the carbocarbenephosphoranes 7 a, 7 b, and 7 c. The last-named molecules have very large first and large second proton affinities. They also bind two BH3 ligands with very high bond energies, which are large enough that the bis-adducts should be isolable in a condensed phase. The second proton affinities of the complexes 5, 6, and 8–10 bearing CO or N2 as ligand are significantly lower than those of the other molecules. However, they give stable complexes with two BH3 ligands and thus are twofold Lewis bases. The calculated data thus identify 1–10 as carbones LCL in which the carbon atom has two electron pairs. The chemistry of carbones is different from that of carbenes because divalent carbon(0) compounds CL2 are π donors and thus may serve as double Lewis bases, while divalent carbon(II) compounds are π acceptors. The theoretical results point toward new directions for experimental research in the field of low-coordinate carbon compounds.
Co-reporter:Andreas Krapp
Theoretical Chemistry Accounts 2010 Volume 127( Issue 3) pp:141-148
Publication Date(Web):2010 October
DOI:10.1007/s00214-009-0696-8
Quantum chemical DFT calculations at the BP86/TZ2P level have been carried out for the complex [HSi(SiH2NH)3Ti–Co(CO)4], which is a model for the experimentally observed compound [MeSi{SiMe2N(4-MeC6H4)}3Ti–Co(CO)4] and for the series of model systems [(H2N)3M–M′(CO)4] (M = Ti, Zr, Hf; M′ = Co, Rh, Ir). The Ti–Co bond in [HSi(SiH2NH)3Ti–Co(CO)4] has a theoretically predicted BDE of De = 59.3 kcal/mol. The bonding analysis suggests that the titanium atom carries a large positive charge, while the cobalt atom is nearly neutral. The covalent and electrostatic contributions to the Ti–Co attraction have similar strength. The Ti–Co bond can be classified as a polar single bond, which has only little π contribution. Calculations of the model compound (H2N)3Ti–Co(CO)4 show that the rotation of the amino groups has a very large influence on the length and on the strength of the Ti–Co bond. The M–M′ bond in the series [(H2N)3M–M′(CO)4] becomes clearly stronger with Ti < Zr < Hf, while the differences between the bond strengths due to change of the atoms M′ are much smaller. The strongest M–M′ bond is predicted for [(H2N)3Hf–Ir(CO)4].
Co-reporter:Mehmet Ali Celik, Robin Haunschild and Gernot Frenking
Organometallics 2010 Volume 29(Issue 7) pp:1560-1568
Publication Date(Web):March 12, 2010
DOI:10.1021/om900986b
Quantum chemical calculations on the hybrid density functional (B3LYP) level of theory were carried out to elucidate the initial step of the catalytic cycle of the aminohydroxylation of olefins with osmium reagents. Various reaction pathways of the model systems (O═)2Os(═NH)2 (Os1) and (O═)2Os(═NH)-cyclo-(−NHCH2CH2HN−) (Os2) and two of their isomers, (HO−)(O═)Os(═NH)(≡N) (Os1′) and (O═)Os(−OH)(≡N)-cyclo-(−NHCH2CH2HN−) (Os2′), with ethylene have been considered. Os1, Os1′, and Os2 prefer a [3+2] cycloaddition kinetically and thermodynamically, whereas Os2′ is predicted to favor a [1+2] pathway due to kinetic reasons. Thermodynamically, Os2′ prefers a consecutive rearrangement after the [1+2] addition. The isomer Os1′ is lower in energy than Os1, but the kinetically and thermodynamically most favorable reaction is the [3+2]NH,NH ethylene addition to the latter isomer, which according to our calculations takes place without a barrier.
Co-reporter:José A. Gámez, Ralf Tonner, and Gernot Frenking
Organometallics 2010 Volume 29(Issue 21) pp:5676-5680
Publication Date(Web):July 28, 2010
DOI:10.1021/om100584e
Quantum chemical calculations using gradient-corrected DFT at the BP86/TZ2P+ level have been carried out for the gallium halide complexes [Fe(CO)4(GaX)] (X = F−I) and for [Fe(CO)5] and [Fe(CO)]4(N2)]. The nature of the metal−ligand bond has been investigated with an energy decomposition analysis. The Fe−GaX bonds in [Fe(CO)4(GaX)] have rather high bond dissociation energies that are smaller than the BDE of [Fe(CO)5] but larger than the BDE of [Fe(CO)]4(N2)]. The axial isomer of [Fe(CO)4(GaF)] is predicted to be slightly more stable than the equatorial form, while the equatorial isomers of the heavier halogen systems [Fe(CO)4(GaX)] (X = Cl−I) are a bit lower in energy than the axial form. The energy difference between the two isomers is for all gallium complexes [Fe(CO)4(GaX)] smaller than 1.0 kcal/mol. The ratio of covalent bonding and electrostatic attraction in the (CO)4Fe−GaX bonds is very similar to the values that are calculated for the (CO)4Fe−CO and (CO)4Fe−N2 bonds. The analysis of the bonding situation reveals that the σ-donation of the ligands GaX is much stronger than the π-back-donation, while the (CO)4Fe−CO and (CO)4Fe−N2 bonds possess equally strong contributions from σ-donation and π-back-donation. The GaX and N2 ligands induce stronger Fe−COtrans bonds in the axial isomers of [Fe(CO)4(GaX)] and [Fe(CO)]4(N2)]: L(CO)3Fe→COtrans π-back-donation.
Co-reporter:Thomas Cadenbach ; Timo Bollermann ; Christian Gemel ; Mustafa Tombul ; Israel Fernandez ; Moritz von Hopffgarten ; Gernot Frenking ;Roland A. Fischer
Journal of the American Chemical Society 2009 Volume 131(Issue 44) pp:16063-16077
Publication Date(Web):October 14, 2009
DOI:10.1021/ja904061w
This paper presents the preparation, characterization and bonding analyses of the closed shell 18 electron compounds [M(ZnR)n] (M = Mo, Ru, Rh, Ni, Pd, Pt, n = 8−12), which feature covalent bonds between n one-electron organo-zinc ligands ZnR (R = Me, Et, η5-C5(CH3)5 = Cp*) and the central metal M. The compounds were obtained in high isolated yields (>80%) by treatment of appropriate GaCp* containing transition metal precursors 13-18, namely [Mo(GaCp*)6], [Ru2(Ga)(GaCp*)7(H)3] or [Ru(GaCp*)6(Cl)2], [(Cp*Ga)4RhGa(η1-Cp*)Me] and [M(GaCp*)4] (M = Ni, Pd, Pt) with ZnMe2 or ZnEt2 in toluene solution at elevated temperatures of 80−110 °C within a few hours of reaction time. Analytical characterization was done by elemental analyses (C, H, Zn, Ga), 1H and 13C NMR spectroscopy. The molecular structures were determined by single crystal X-ray diffraction. The coordination environment of the central metal M and the M−Zn and Zn−Zn distances mimic the situation in known solid state M/Zn Hume−Rothery phases. DFT calculations at the RI-BP86/def2-TZVPP and BP86/TZ2P+ levels of theory, AIM and EDA analyses were done with [M(ZnH)n] (M = Mo, Ru, Rh, Pd; n = 12, 10, 9, 8) as models of the homologous series. The results reveal that the molecules can be compared to 18 electron gold clusters of the type M@Aun, that is, W@Au12, but are neither genuine coordination compounds nor interstitial cage clusters. The molecules are held together by strong radial M−Zn bonds. The tangential Zn−Zn interactions are generally very weak and the (ZnH)n cages are not stable without the central metal M.
Co-reporter:Israel Fernández ; C. Adam Dyker ; Alan DeHope ; Bruno Donnadieu ; Gernot Frenking ;Guy Bertrand
Journal of the American Chemical Society 2009 Volume 131(Issue 33) pp:11875-11881
Publication Date(Web):July 31, 2009
DOI:10.1021/ja903396e
Small ring allenes are usually highly strained and highly reactive species, and for a long time considered only as transient intermediates. The recent isolation of a five membered heterocyclic allene 1f has raised questions and debate regarding the factors responsible for its stability. Since 1f has been derived by deprotonation of a pyrazolium ion 2f, it has been suggested that the stability of 1f comes from its aromatic character. Here we report computational evidence, including HOMA and NICS aromaticity indices, that allenes derived from 3,5-bis(π-donor) substituted pyrazolium salts are weakly aromatic to nonaromatic, and that even their pyrazolium ion precursors have dramatically reduced aromaticity. Exocyclic delocalization, involving the π-donor substituents, occurs at the expense of aromaticity and increases with the strength of the donor. Experimental support for these conclusions is found in the crystallographically determined structure of 3,5-bis(dimethlamino)pyrazolium ion 2g, which exhibits highly pyramidalized endocyclic nitrogen centers but planarized exocyclic ones, and from the facile C4 protonation to give a stable pyrazole-1,2-diium salt 3g, which has also been crystallographically characterized.
Co-reporter:Gernot Frenking
Journal of the American Chemical Society 2009 Volume 131(Issue 25) pp:8989-8999
Publication Date(Web):May 29, 2009
DOI:10.1021/ja902198z
Room temperature photolysis of aminoborylene complexes, [(CO)5M═B═N(SiMe3)2] (1: M = Cr, 2: Mo) in the presence of a series of alkynes and diynes, 1,2-bis(4-methoxyphenyl)ethyne, 1,2-bis(4-(trifluoromethyl)phenyl)ethyne, 1,4-diphenylbuta-1,3-diyne, 1,4-bis(4-methoxyphenyl)buta-1,3-diyne, 1,4-bis(trimethylsilylethynyl)benzene and 2,5-bis(4-N,N-dimethylaminophenylethynyl)thiophene led to the isolation of novel mono and bis-bis-(trimethylsilyl)aminoborirenes in high yields, that is [(RC═CR)(μ-BN(SiMe3)2], (3: R = C6H4-4-OMe and 4: R = C6H4-4-CF3); [{(μ-BN(SiMe3)2(RC═C−)}2], (5: R = C6H5 and 6: R = C6H4-4-OMe); [1,4-bis-{(μ-BN(SiMe3)2(SiMe3C═C)}benzene], 7 and [2,5-bis-{(μ-BN(SiMe3)2 ((C6H4NMe2)C═C)}-thiophene], 8. All borirenes were isolated as light yellow, air and moisture sensitive solids. The new borirenes have been characterized in solution by 1H, 11B, 13C NMR spectroscopy and elemental analysis and the structural types were unequivocally established by crystallographic analysis of compounds 6 and 7. DFT calculations were performed to evaluate the extent of π-conjugation between the electrons of the carbon backbone and the empty pz orbital of the boron atom, and TD-DFT calculations were carried out to examine the nature of the electronic transitions.
Co-reporter:Greta Heydenrych;Moritz von Hopffgarten;Elzet Ster;Oliver Schuster;Helgard G. Raubenheimer
European Journal of Inorganic Chemistry 2009 Volume 2009( Issue 13) pp:1892-1904
Publication Date(Web):
DOI:10.1002/ejic.200801244
Abstract
Quantum chemical calculations using DFT at the BP86/TZ2P level of theory are reported for the complexes (PH3)2ClM-L where L is an N-heterocyclic ligand and M a group-10 metal Ni, Pd and Pt. The ligands comprise pyridyl groups or carbenes derived from the pyridine, quinolidine or isoquinolidine systems wherein the nitrogen atom is either adjacent to the carbene carbon atom or it is in a remote (meta or para, or in the adjacent ring) position. Comparative calculations include the isomeric ligands of the well-known five-membered N-heterocyclic carbene. The nature of the metal–ligand interactions are investigated by energy decomposition analysis (EDA). The EDA results suggest that the nature of the metal–carbene bonds in the complexes shows little variation when the position of the nitrogen atom in pyridylidenes is adjacent (ortho) or remote (meta or para). It changes even very little when the nitrogen atom is in an adjacent ring to the cyclic carbene moiety. The most significant differences between the bond strengths come from the energy level of the σ-HOMO of the carbene ligand which depends largely on the position of the nitrogen atom. An energetically higher-lying σ lone-pair orbital of the carbene ligand yields stronger orbital interactions but also stronger electrostatic attraction because of better overlap with the metal nucleus. This holds also for the isomers of the five-membered N-heterocyclic carbenes. An excellent correlation is established between the ϵ(HOMO) values of the ligands and the metal–ligand interaction energies, ΔEint.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2009)
Co-reporter:C.ÓscarC. Jiménez-Halla;Israel Fernández Dr. Dr.
Angewandte Chemie International Edition 2009 Volume 48( Issue 2) pp:366-369
Publication Date(Web):
DOI:10.1002/anie.200803252
Co-reporter:Robin Haunschild, Sandor Tüllmann, Gernot Frenking, Max C. Holthausen
Journal of Organometallic Chemistry 2009 694(7–8) pp: 1081-1090
Publication Date(Web):
DOI:10.1016/j.jorganchem.2008.10.017
Co-reporter:Holger Braunschweig Dr.;Thomas Herbst Dr.;Krzysztof Radacki Dr. Dr.;MehmetAli Celik
Chemistry - A European Journal 2009 Volume 15( Issue 44) pp:12099-12106
Publication Date(Web):
DOI:10.1002/chem.200901749
Abstract
We report on the hydroboration of 1-[bis(trimethylsilyl)amino]-2,3-diethylborirene (3) with 9-borabicyclo[3.3.1]nonane (9-BBN), which led through ring-opening to an amino(vinyl)borane. The viscous borane was subsequently converted into a crystalline borate on treatment with MeLi. Both compounds were fully characterized by multinuclear NMR spectroscopy and in case of the latter by single-crystal X-ray diffraction analysis. To elucidate the reaction mechanism of the unexpected boron-carbon bond cleavage, DFT calculations of energy minima and transition states for the hydroboration were carried out.
Co-reporter:Nozomi Takagi Dr.;Takayasu Shimizu Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 34) pp:8593-8604
Publication Date(Web):
DOI:10.1002/chem.200901401
Abstract
Quantum-chemical calculations at the BP86/TZVPP level of theory have been carried out for compounds EL2 for E=Si, Ge, Sn, where L is a five-membered cyclic ylidene or N-heterocyclic ylidene. The theoretical results provide evidence for the classification of the complexes as divalent E(0) compounds, where the bonding situation is best described in terms of donor–acceptor interactions between a bare atom E, which retains its valence electrons as two lone pairs, and two donor ligands LEL. The molecules are very strong donors, which may bind one or two Lewis acids. Divalent E(0) compounds have unusually high second proton affinities and they are strong σ donor ligands. The calculations predict that complexes of EL2 with one or two BH3 ligands are stable enough to become isolated in a condensed phase. It is also shown that the bond dissociation energies (BDEs) of transition-metal complexes [(CO)5WD] and [(CO)3NiD], where D=EL2 are rather high. The BDE of some ligands D are higher than those of CO in the metal carbonyls.
Co-reporter:Nozomi Takagi Dr.;Takayasu Shimizu Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 14) pp:3448-3456
Publication Date(Web):
DOI:10.1002/chem.200802739
Co-reporter:Anastas Sidiropoulos;Cameron Jones ;Andreas Stasch Dr.;Susanne Klein Dr.
Angewandte Chemie International Edition 2009 Volume 48( Issue 51) pp:9701-9704
Publication Date(Web):
DOI:10.1002/anie.200905495
Co-reporter:Moh Melaimi Dr.;Pattiyil Parameswaran Dr.;Bruno Donnadieu ;Guy Bertr
Angewandte Chemie 2009 Volume 121( Issue 26) pp:4886-4889
Publication Date(Web):
DOI:10.1002/ange.200901117
Co-reporter:Moh Melaimi Dr.;Pattiyil Parameswaran Dr.;Bruno Donnadieu ;Guy Bertr
Angewandte Chemie International Edition 2009 Volume 48( Issue 26) pp:4792-4795
Publication Date(Web):
DOI:10.1002/anie.200901117
Co-reporter:Pattiyil Parameswaran Dr. Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 35) pp:8817-8824
Publication Date(Web):
DOI:10.1002/chem.200900792
Abstract
Density functional calculations at the BP86/TZ2P level were carried out to understand the ligand properties of the 16-valence-electron(VE) Group 14 complexes [(PMe3)2Cl2M(E)] (1ME) and the 18-VE Group 14 complexes [(PMe3)2(CO)2M(E)] (2ME; M=Fe, Ru, Os; E=C, Si, Ge, Sn) in complexation with W(CO)5. Calculations were also carried out for the complexes (CO)5W–EO. The complexes [(PMe3)2Cl2M(E)] and [(PMe3)2(CO)2M(E)] bind strongly to W(CO)5 yielding the adducts 1ME–W(CO)5 and 2ME–W(CO)5, which have C2v equilibrium geometries. The bond strengths of the heavier Group 14 ligands 1ME (E=Si–Sn) are uniformly larger, by about 6–7 kcal mol−1, than those of the respective EO ligand in (CO)5W-EO, while the carbon complexes 1MC–W(CO)5 have comparable bond dissociation energies (BDE) to CO. The heavier 18-VE ligands 2ME (E=Si–Sn) are about 23–25 kcal mol−1 more strongly bonded than the associated EO ligand, while the BDE of 2MC is about 17–21 kcal mol−1 larger than that of CO. Analysis of the bonding with an energy-decomposition scheme reveals that 1ME is isolobal with EO and that the nature of the bonding in 1ME–W(CO)5 is very similar to that in (CO)5W–EO. The ligands 1ME are slightly weaker π acceptors than EO while the π-acceptor strength of 2ME is even lower.
Co-reporter:Pattiyil Parameswaran Dr. Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 35) pp:8807-8816
Publication Date(Web):
DOI:10.1002/chem.200900791
Abstract
The equilibrium geometries and bond dissociation energies of 16-valence-electron(VE) complexes [(PMe3)2Cl2M(E)] and 18-VE complexes [(PMe3)2(CO)2M(E)] with M=Fe, Ru, Os and E=C, Si, Ge, Sn were calculated by using density functional theory at the BP86/TZ2P level. The nature of the ME bond was analyzed with the NBO charge decomposition analysis and the EDA energy-decomposition analysis. The theoretical results predict that the heavier Group 14 complexes [(PMe3)2Cl2M(E)] and [(PMe3)2(CO)2M(E)] with E=Si, Ge, Sn have C2v equilibrium geometries in which the PMe3 ligands are in the axial positions. The complexes have strong ME bonds which are slightly stronger in the 16-VE species 1ME than in the 18-VE complexes 2ME. The calculated bond dissociation energies show that the ME bonds become weaker in both series in the order C>Si>Ge>Sn; the bond strength increases in the order Fe<Ru<Os for 1ME, whereas a U-shaped trend Ru<Os<Fe is found for 2ME. The ME bonding analysis suggests that the 16-VE complexes 1ME have two electron-sharing bonds with σ and π symmetry and one donor–acceptor π bond like the carbon complex. Thus, the bonding situation is intermediate between a typical Fischer complex and a Schrock complex. In contrast, the 18-VE complexes 2ME have donor–acceptor bonds, as suggested by the Dewar–Chatt–Duncanson model, with one ME σ donor bond and two ME π-acceptor bonds, which are not degenerate. The shape of the frontier orbitals reveals that the HOMO−2 σ MO and the LUMO and LUMO+1 π* MOs of 1ME are very similar to the frontier orbitals of CO.
Co-reporter:Robin Haunschild, Gernot Frenking
Journal of Organometallic Chemistry 2009 694(25) pp: 4090-4093
Publication Date(Web):
DOI:10.1016/j.jorganchem.2009.08.022
Co-reporter:C.ÓscarC. Jiménez-Halla;Israel Fernández Dr. Dr.
Angewandte Chemie 2009 Volume 121( Issue 2) pp:372-375
Publication Date(Web):
DOI:10.1002/ange.200803252
Co-reporter:Ralf Tonner and Gernot Frenking
Organometallics 2009 Volume 28(Issue 13) pp:3901-3905
Publication Date(Web):June 8, 2009
DOI:10.1021/om900206w
The complexes [D-Ni(CO)3] and [D-RhCl(CO)2] have been calculated with DFT methods where D is a strong carbon or phosphorus donor ligand. The focus of the work is on the estimate of the donor strength of divalent carbon(0) compounds CL2 compared to common carbon(II) ligands (N-heterocyclic carbenes, NHC), phosphines, and several ligands recently introduced. Tolman’s electronic parameters are derived by a fit procedure to experimental data to enable better comparison of different donor strength scales. It turns out that carbon(0) ligands CL2 with L = PR3 (carbodiphosphoranes) or NHC (carbodicarbenes) are much stronger donors than carbene and phosphine ligands. Besides the calculation of known substitution patterns, several new ligands of this class are proposed that might be valuable in the search for new ligands with adjustable donor properties.
Co-reporter:Ralf Tonner and Gernot Frenking
Chemical Communications 2008 (Issue 13) pp:1584-1586
Publication Date(Web):27 Feb 2008
DOI:10.1039/B717511F
Theoretical investigations suggest that substitution of an N-heterocyclic carbene by a carbodiphosphorane in the Grubb’s catalyst for olefin metathesis might lead to enhanced reactivity.
Co-reporter:Milind M. Deshmukh, Shridhar R. Gadre, Ralf Tonner and Gernot Frenking
Physical Chemistry Chemical Physics 2008 vol. 10(Issue 17) pp:2298-2301
Publication Date(Web):10 Mar 2008
DOI:10.1039/B803068E
The molecular electrostatic potentials of divalent carbon(0) and divalent carbon(II) compounds are calculated and the results are compared with theoretically predicted proton affinities and complexation energies with BH3.
Co-reporter:Israel Fernández, Erick Cerpa, Gabriel Merino and Gernot Frenking
Organometallics 2008 Volume 27(Issue 6) pp:1106-1111
Publication Date(Web):February 22, 2008
DOI:10.1021/om700994y
The geometries and metal–ligand bond dissociation energies of [E−Cp−E′]+ complexes (E, E′ = group 13 element; Cp = cyclopentadienyl) have been calculated within the density functional theory framework. The geometries of the title complexes were optimized at the BP86 level with the TZ2P valence basis set. The nature of the metal–ligand bonding has been studied using the energy decomposition analysis (EDA). The calculated bond strengths for the homoleptic complexes [E−Cp−E]+ with respect to loss of a neutral or charged group 13 atom are Ga > Al > In > Tl ≫ B. While the energetically most favorable pathway for the boron complex [B−Cp−B]+ is the loss of a neutral boron atom, heavier homologues [E−Cp−E]+ (E = Al−Tl) dissociate via loss of the charged atom E+. The heteroleptic species [E−Cp−E′]+ are less stable than the homoleptic complexes [E−Cp−E]+. The lowest energy pathway for dissociation is the loss of the positively charged heavier atom E′+. The B−Cp interactions in the boron complexes have a larger (covalent) orbital character than the E−Cp bonding in the heavier homologues. The energy decomposition analysis of [E−Cp−E′]+, using Cp− and (E···E′)2+ as ligands, suggests that the a1(σ) bonding has nearly the same strength as the e1(π) bonding.
Co-reporter:Ralf Tonner Dipl.-Chem. Dr.
Chemistry - A European Journal 2008 Volume 14( Issue 11) pp:3260-3272
Publication Date(Web):
DOI:10.1002/chem.200701390
Abstract
Quantum-chemical calculations with DFT (BP86) and ab initio methods [MP2, SCS-MP2, CCSD(T)] have been carried out for the molecules C(PH3)2 (1), C(PMe3)2 (2), C(PPh3)2 (3), C(PPh3)(CO) (4), C(CO)2 (5), C(NHCH)2 (6), C(NHCMe)2 (7) (Me2N)2CCC(NMe2)2 (8), and NHC (9), where NHC=N-heterocyclic carbene and NHCMe=N-methyl-substituted NHC. The electronic structure in 1–9 was analyzed with charge- and energy-partitioning methods. The results show that the bonding situations in L2C compounds 1–8 can be interpreted in terms of donor–acceptor interactions between closed-shell ligands L and a carbon atom which has two lone-pair orbitals LCL. This holds particularly for the carbodiphosphoranes 1–3 where L=PR3, which therefore are classified as divalent carbon(0) compounds. The NBO analysis suggests that the best Lewis structures for the carbodicarbenes 6 and 7 where L is a NHC ligand have CCC double bonds as in the tetraaminoallene 8. However, the Lewis structures of 6–8, in which two lone-pair orbitals at the central carbon atom are enforced, have only a slightly higher residual density. Visual inspection of the frontier orbitals of the latter species reveals their pronounced lone-pair character, which suggests that even the quasi-linear tetraaminoallene 8 is a “masked” divalent carbon(0) compound. This explains the very shallow bending potential of 8. The same conclusion is drawn for phosphoranylketene 4 and for carbon suboxide (5), which according to the bonding analysis have hidden double-lone-pair character. The AIM analysis and the EDA calculations support the assignment of carbodiphosphoranes as divalent carbon(0) compounds, while NHC 9 is characterized as a divalent carbon(II) compound. The LC(1D) donor–acceptor bonds are roughly twice as strong as the respective LBH3 bond.
Co-reporter:Moritz vonHopffgarten Dipl.-Chem. Dr.
Chemistry - A European Journal 2008 Volume 14( Issue 33) pp:10227-10231
Publication Date(Web):
DOI:10.1002/chem.200801351
Co-reporter:Gaddamanugu Gayatri, Yarasi Soujanya, Israel Fernández, Gernot Frenking and G. Narahari Sastry
The Journal of Physical Chemistry A 2008 Volume 112(Issue 50) pp:12919-12924
Publication Date(Web):June 13, 2008
DOI:10.1021/jp802886g
The computational study explores the electronic fine tuning of the exocyclic C−C single bond length in tetrahedranyl tetrahedrane as a function of various substituents. The factors which determine the bond lengths and bond strengths are examined by using the EDA method.
Co-reporter:Ralf Tonner Dr.;Greta Heydenrych Dr. Dr.
ChemPhysChem 2008 Volume 9( Issue 10) pp:1474-1481
Publication Date(Web):
DOI:10.1002/cphc.200800208
Abstract
Quantum chemical calculations at the MP2/TZVPP//BP86/SVP level are reported for the first and second proton affinities (PAs) of divalent carbon-donor molecules. The molecules investigated are imidazol-2-ylidenes (“normal” NHCs) and the tautomeric imidazol-4/5-ylidenes (“abnormal” NHCs). PAs are also calculated for acyclic and cyclic carbodiphosphoranes, carbophosphoranesulfide, unsaturated and saturated carbodicarbenes, tetraaminoallenes and carbon suboxide. The results are discussed in terms of divalent carbon(II) compounds (carbenes) CR2, which have one lone electron pair at carbon, and carbon(0) compounds CL2, which have two lone pairs at carbon and two CL donor–acceptor bonds. Divalent C(0) compounds (carbones) not only have very high first PAs, but the second PA is also large and strong enough to isolate doubly protonated C(0) species as salts in a condensed phase. The first PA of divalent carbon(II) compounds (carbenes) are also large. However, they have much smaller second PAs than the divalent carbon(0) compounds. The divalent C(0) character of a compound is not always obvious when the bonding situation in the equilibrium geometry is considered. This is the case, for example, for tetraaminoallenes (TAAs). Protonation of TAAs changes the bonding situation of the central moiety from doubly bonded (R2N)2CCC(NR2)2 to a donor–acceptor description (R2N)2CC(H+)nC(NR2) [n=1, 2]. The atomic partial charge at the carbon donor atom does not correlate with the PA and the trend of the second PA may be quite different from the trend of the first. The trends of the first and second PA correlates quite well with the eigenvalues of the highest-lying carbon lone-pair orbitals.
Co-reporter:Andreas Krapp Dr. Dr.;Einar Uggerud Dr.
Chemistry - A European Journal 2008 Volume 14( Issue 13) pp:4028-4038
Publication Date(Web):
DOI:10.1002/chem.200701613
Abstract
The electronic structures and bonding patterns for a new class of radical cations, [HnE-H-H-EHn]+ (EHn=element hydride, E=element of Groups 15–18), have been investigated by applying quantum-chemical methods. All structures investigated give rise to symmetric potential energy minimum structures. We envisage clear periodic trends. The HH bond length is shorter for elements toward the bottom of the periodic table of elements, and a short HH bond corresponds to accumulation of electron density in the central HH region. All [HnE-H-H-EHn]+ of Groups 15–17 are thermodynamically unstable towards loss of either H2 or H. The barriers for these dissociations are rather low. The Group 18 congeners, except E=Xe, appear to be global minima of the respective potential energy surfaces. The findings are discussed in terms of H2 bond activation, and a general mechanistic scheme for the standard reduction process 2H+ + 2e− H2 is given. Finally, it is proposed that some of the symmetric radical cations are likely to be observed in mass spectrometric or matrix isolation experiments.
Co-reporter:Ralf Tonner Dipl.-Chem. Dr.
Chemistry - A European Journal 2008 Volume 14( Issue 11) pp:3273-3289
Publication Date(Web):
DOI:10.1002/chem.200701392
Abstract
Quantum-chemical calculations with DFT (BP86) and ab initio methods (MP2, SCS-MP2) were carried out for protonated and diprotonated compounds N-H+ and N-(H+)2 and for the complexes N-BH3, N-(BH3)2, N-CO2, N-(CO2)2, N-W(CO)5, N-Ni(CO)3 and N-Ni(CO)2 where N=C(PH3)2 (1), C(PMe3)2 (2), C(PPh3)2 (3), C(PPh3)(CO) (4), C(CO)2 (5), C(NHCH)2 (6), C(NHCMe)2 (7) (Me2N)2CCC(NMe2)2 (8) and NHC (9) (NHCH=N-heterocyclic carbene, NHCMe=N-substituted N-heterocyclic carbene). Compounds 1–4 and 6–9 are very strong electron donors, and this is manifested in calculated protonation energies that reach values of up to 300 kcal mol−1 for 7 and in very high bond strengths of the donor–acceptor complexes. The electronic structure of the compounds was analyzed with charge- and energy-partitioning methods. The calculations show that the experimentally known compounds 2–5 and 8 chemically behave like molecules L2C which have two LC donor–acceptor bonds and a carbon atom with two electron lone pairs. The behavior is not directly obvious when the linear structures of carbon suboxide and tetraaminoallenes are considered. They only come to the fore on reaction with strong electron-pair acceptors. The calculations predict that single and double protonation of 5 and 8 take place at the central carbon atom, where the negative charge increases upon subsequent protonation. The hitherto experimentally unknown carbodicarbenes 6 and 7 are predicted to be even stronger Lewis bases than the carbodiphosphoranes 1–3.
Co-reporter:Thomas Cadenbach;Timo Bollermann;Christian Gemel Dr.;Israel Fernez Dr.;Moritz vonHopffgarten Dr.;RolA. Fischer Dr.
Angewandte Chemie 2008 Volume 120( Issue 47) pp:9290-9295
Publication Date(Web):
DOI:10.1002/ange.200802811
Co-reporter:Andreas Krapp Dr.
Angewandte Chemie 2008 Volume 120( Issue 41) pp:7912-7913
Publication Date(Web):
DOI:10.1002/ange.200705423
Co-reporter:Gernot Frenking Dr.
Angewandte Chemie 2008 Volume 120( Issue 38) pp:7280-7281
Publication Date(Web):
DOI:10.1002/ange.200802500
No abstract is available for this article.
Co-reporter:Andreas Krapp Dr.
Angewandte Chemie International Edition 2008 Volume 47( Issue 41) pp:7796-7797
Publication Date(Web):
DOI:10.1002/anie.200705423
Co-reporter:Holger Braunschweig Dr.;Israel Fernández Dr. Dr.;Thomas Kupfer Dr.
Angewandte Chemie International Edition 2008 Volume 47( Issue 10) pp:1951-1954
Publication Date(Web):
DOI:10.1002/anie.200704771
Co-reporter:Gernot Frenking Dr.
Angewandte Chemie International Edition 2008 Volume 47( Issue 38) pp:7168-7169
Publication Date(Web):
DOI:10.1002/anie.200802500
No abstract is available for this article.
Co-reporter:Thomas Cadenbach;Timo Bollermann;Christian Gemel Dr.;Israel Fernez Dr.;Moritz vonHopffgarten Dr.;RolA. Fischer Dr.
Angewandte Chemie International Edition 2008 Volume 47( Issue 47) pp:9150-9154
Publication Date(Web):
DOI:10.1002/anie.200802811
Co-reporter:Thomas Cadenbach;Timo Bollermann;Christian Gemel Dr.;Israel Fernez Dr.;Moritz vonHopffgarten Dr.;RolA. Fischer Dr.
Angewandte Chemie International Edition 2008 Volume 47( Issue 47) pp:
Publication Date(Web):
DOI:10.1002/anie.200890238
Co-reporter:Thomas Cadenbach;Timo Bollermann;Christian Gemel Dr.;Israel Fernez Dr.;Moritz vonHopffgarten Dr.;RolA. Fischer Dr.
Angewandte Chemie 2008 Volume 120( Issue 47) pp:
Publication Date(Web):
DOI:10.1002/ange.200890287
Co-reporter:Giovanni F. Caramori
Theoretical Chemistry Accounts 2008 Volume 120( Issue 4-6) pp:351-361
Publication Date(Web):2008 July
DOI:10.1007/s00214-008-0435-6
Quantum chemical calculations at the DFT level have been carried out for model complexes [Mo(P)(NH2)3] (1), [Mo(N)(NH2)3] (2), [Mo(PO)(NH2)3] (3), [Mo(NO)(NH2)3] (4), [Mo(CO)5(PO)]+ (5), and [Mo(CO)5(NO)]+ (6). The equilibrium geometries and the vibration frequencies are in good agreement with experimental and previous theoretical results. The nature of the Mo–PO, Mo–NO, Mo–PO+, Mo–NO+, Mo–P, and Mo–N bond has been investigated by means of the AIM, NBO and EDA methods. The NBO and EDA data complement each other in the interpretation of the interatomic interactions while the numerical AIM results must be interpreted with caution. The terminal Mo–P and Mo–N bonds in 1 and 2 are clearly electron-sharing triple bonds. The terminal Mo–PO and Mo–NO bonds in 3 and 4 have also three bonding contributions from a σ and a degenerate π orbital where the σ components are more polarized toward the ligand end and the π orbitals are more polarized toward the metal end than in 1 and 2. The EDA calculations show that the π bonding contributions to the Mo–PO and Mo–NO bonds in 3 and 4 are much more important than the σ contributions while σ and π bonding have nearly equal strength in the terminal Mo–P and Mo–N bonds in 1 and 2. The total (NH2)3Mo–PO binding interactions are stronger than for (NH2)3Mo–P which is in agreement with the shorter Mo–PO bond. The calculated bond orders suggest that there are only (NH2)3Mo–PO and (NH2)3Mo–NO double bonds which comes from the larger polarization of the σ and π contributions but a closer inspection of the bonding shows that these bonds should also be considered as electron-sharing triple bonds. The bonding situation in the positively charged complexes [(CO)5Mo–(PO)]+ and [(CO)5Mo–(NO)]+ is best described in terms of (CO)5Mo → XO+ donation and (CO)5Mo ← XO+ backdonation (X = P, N) using the Dewar–Chatt–Duncanson model. The latter bonds are stronger and have a larger π character than the Mo-CO bonds.
Co-reporter:Robin Haunschild;Christoph Loschen;Sor Tüllmann;Daniel Cappel;Max C. Holthausen;Markus Hölscher
Journal of Physical Organic Chemistry 2007 Volume 20(Issue 1) pp:11-18
Publication Date(Web):9 FEB 2007
DOI:10.1002/poc.1095
Quantum chemical calculations using density functional theory at the B3LYP level in combination with relativistic effective core potentials for the metals and TZ2P valence basis sets have been carried out for elucidating the reaction pathways of ethylene addition to MeReO2(CH2) (C1). The results are compared with our previous studies of ethylene addition to OsO2(CH2)2 (A1) and OsO3(CH2) (B1). Significant differences have been found between the ethylene additions to the osmium compounds A1 and B1 and the rhenium compound C1. Seven pathways for the reaction C1+C2H4 were studied, but only the [2+2]Re,C addition yielding rhenacyclobutane C5 is an exothermic process with a high activation barrier of 48.9 kcal mol−1. The lowest activation energy (27.7 kcal mol−1) is calculated for the [2+2]Re,C addition, which leads to the isomeric form C5′. Two further concerted reactions [3+2]O,C, [3+2]O,O, and [2+2]Re,O and the addition/hydrogen migration of ethylene to one oxo ligand are endothermic processes which have rather high activation barriers (>35 kcal mol−1). Four isomerization processes of C1 have very large activation energies of >65 kcal mol−1. The ethylene addition to the osmium compounds A1 and B1 are much more exothermic and have lower activation barriers than the C2H4 addition to C1. Copyright © 2007 John Wiley & Sons, Ltd.
Co-reporter:Gernot Frenking Dr.;Bernhard Neumüller Dr.;Wolfgang Petz Dr.;Ralf Tonner Dipl. Chem.;Florian Öxler Dipl. Chem.
Angewandte Chemie International Edition 2007 Volume 46(Issue 17) pp:
Publication Date(Web):1 MAR 2007
DOI:10.1002/anie.200700327
Neither is the synthesis of a molecule a mere technical conformation of a theoretical prediction, nor is a quantum chemical investigation a mere supplement to experimental work—the two sides rather complement each other. The theoretical study in question has shown that more compounds of the general formula EL2 with donor–acceptor bonds may exist which await synthesis.
Co-reporter:Holger Braunschweig Dr.;Israel Fernández Dr. Dr.;Krzysztof Radacki Dr.;Fabian Seeler Dipl.-Chem.
Angewandte Chemie International Edition 2007 Volume 46(Issue 27) pp:
Publication Date(Web):30 MAY 2007
DOI:10.1002/anie.200700382
The photochemically induced transfer of a ferroborylene yielded the first metal-bound borirene (see scheme and product structure, in which the methyl substituents are omitted). Experimental and theoretical data suggest significant π delocalization in the three-membered BC2 ring.
Co-reporter:Ralf Tonner;Florian Öxler;Bernhard Neumüller Dr.;Wolfgang Petz Dr. Dr.
Angewandte Chemie International Edition 2007 Volume 46(Issue 28) pp:
Publication Date(Web):2 JUL 2007
DOI:10.1002/anie.200790136
Co-reporter:Holger Braunschweig Dr.;Israel Fernández Dr. Dr.;Krzysztof Radacki Dr.;Fabian Seeler Dipl.-Chem.
Angewandte Chemie 2007 Volume 119(Issue 27) pp:
Publication Date(Web):30 MAY 2007
DOI:10.1002/ange.200700382
Der photochemisch induzierte Transfer eines Ferroborylens lieferte das erste metallgebundene Boriren (siehe Schema und Struktur des Produkts ohne Methylsubstituenten). Experimentelle und theoretische Daten belegen eine deutliche π-Delokalisierung innerhalb des BC2-Dreirings.
Co-reporter:Gernot Frenking Dr.;Bernhard Neumüller Dr.;Wolfgang Petz Dr.;Ralf Tonner Dipl. Chem.;Florian Öxler Dipl. Chem.
Angewandte Chemie 2007 Volume 119(Issue 17) pp:
Publication Date(Web):1 MAR 2007
DOI:10.1002/ange.200700327
Weder ist die Synthese eines Moleküls eine bloße technische Bestätigung einer theoretischen Vorhersage noch sind quantenchemische Untersuchungen eine bloße Ergänzung experimenteller Arbeiten – vielmehr sind beide Seiten komplementär. Die vorgestellte Studie zu Carbodiphosphoranen hat gezeigt, dass noch mehr Verbindungen mit der allgemeinen Formel EL2 mit Donor-Acceptor-Bindungen existieren sollten, deren Synthese in Aussicht steht.
Co-reporter:Ralf Tonner;Florian Öxler;Bernhard Neumüller Dr.;Wolfgang Petz Dr. Dr.
Angewandte Chemie 2007 Volume 119(Issue 28) pp:
Publication Date(Web):2 JUL 2007
DOI:10.1002/ange.200790136
Co-reporter:Ralf Tonner Dr.
Angewandte Chemie 2007 Volume 119(Issue 45) pp:
Publication Date(Web):8 OCT 2007
DOI:10.1002/ange.200701632
Quantenchemische Rechnungen liefern eindeutige Belege, dass die experimentell noch unbekannten Carbodicarbene C(NHC)2 (NHC=N-heterocyclischer Carbenligand; Beispiel siehe Bild) eine präparativ zugängliche Klasse von Molekülen mit zweibindigem Kohlenstoff(0) sind, die als sehr starke Nucleophile und sehr starke Basen wirken und interessante Eigenschaften als Liganden in Übergangsmetallkomplexen aufweisen sollten.
Co-reporter:Ralf Tonner Dr.
Angewandte Chemie International Edition 2007 Volume 46(Issue 45) pp:
Publication Date(Web):8 OCT 2007
DOI:10.1002/anie.200701632
Future targets: Quantum-chemical calculations predict that the experimentally still unknown carbodicarbenes C(NHC)2 (NHC=N-heterocyclic carbene; see picture for example) are a synthetically accessible class of divalent carbon(0) compounds which are very strong nucleophiles and bases that may be useful ligands for transition-metal complexes.
Co-reporter:Israel Fernández Dr.;Einar Uggerud
Chemistry - A European Journal 2007 Volume 13(Issue 30) pp:
Publication Date(Web):23 JUL 2007
DOI:10.1002/chem.200700744
We report that only elements more electropositive than carbon (Group 13, 14, and Be) form stable symmetrical HnECH3EHn+ structures (E=Group 1, 2, 13, or 14 element) with a planar CH3 group symmetrically bonded to two EHn moieties, in analogy with prototypical SN2 transition structures. Analysis of the bonding situation of these pentacoordinate carbon molecules was studied by means of an energy decomposition analysis (EDA) of the interaction energy. This shows that HnECH3EHn+ molecules can be viewed as being composed of one CH3 group that is σ-covalently bonded to two EHn groups forming a three-center, two-electron bond.
Co-reporter:Andreas Krapp Dipl.-Chem.
Chemistry - A European Journal 2007 Volume 13(Issue 29) pp:
Publication Date(Web):18 JUL 2007
DOI:10.1002/chem.200700467
Quantum-chemical calculations using DFT (BP86) and ab initio methods (MP2, SCS-MP2) have been carried out for the endohedral fullerenes Ng2@C60 (Ng=He–Xe). The nature of the interactions has been analyzed with charge- and energy-partitioning methods and with the topological analysis of the electron density (Atoms-in-Molecules (AIM)). The calculations predict that the equilibrium geometries of Ng2@C60 have D3d symmetry when Ng=Ne, Ar, Kr, while the energy-minimum structure of Xe2@C60 has D5d symmetry. The precession movement of He2 in He2@C60 has practically no barrier. The NgNg distances in Ng2@C60 are much shorter than in free Ng2. All compounds Ng2@C60 are thermodynamically unstable towards loss of the noble gas atoms. The heavier species Ar2@C60, Kr2@C60, and Xe2@C60 are high energy compounds which are at the BSSE corrected SCS-MP2/TZVPP level in the range 96.7–305.5 kcal mol−1 less stable than free C60 + 2 Ng. The AIM method reveals that there is always an NgNg bond path in Ng2@C60. There are six NgC bond paths in (D3d) Ar2@C60, Kr2@C60, and Xe2@C60, whereas the lighter D3d homologues He2@C60 and Ne2@C60 have only three NgC2 paths. The calculated charge distribution and the orbital analysis clearly show that the bonding situation in Xe2@C60 significantly differs from those of the lighter homologues. The atomic partial charge of the [Xe2] moiety is +1.06, whereas the charges of the lighter dimers [Ng2] are close to zero. The a2u HOMO of (D3d) Xe2@C60 in the 1A1g state shows a large mixing of the highest lying occupied σ* orbital of [Xe2] and the orbitals of the C60 cage. There is only a small gap between the a2u HOMO of Xe2@C60 and the eu LUMO and the a2u LUMO+1. The calculations show that there are several triplet states which are close in energy to each other and to the 1A1g state. The bonding analysis suggests that the interacting species in Xe2@C60 are the charged species Xe2q+ and C60q−, where 1<q<2. The calculated XeXe distance in the endohedral fullerene (2.494 Å) is even shorter than the calculated value for free Xe22+ (2.746 Å). Thus, the XeC and XeXe interactions in Xe2@C60 should be considered as genuine chemical bonds which are enforced by the compression energy. The NgNg and NgC interactions in the lighter homologues Ar2@C60 and Kr2@C60 may also be considered as chemical bonds because the theoretically predicted properties of the endohedral fullerenes are significantly different from the free C60 and noble gas atoms. According to the bonding analysis, He2@C60 and Ne2@C60 are weakly bonded van der Waals complexes.
Co-reporter:Israel Fernández Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 20) pp:
Publication Date(Web):23 APR 2007
DOI:10.1002/chem.200601674
The electronic structure and bonding situation in 21 metallabenzenes (metal=Os, Ru, Ir, Rh, Pt, and Pd) were investigated at the DFT level (BP86/TZ2P) by using an energy decomposition analysis (EDA) of the interaction energy between various fragments. The aim of the work is to estimate the strength of the π bonding and the aromatic character of the metallacyclic compounds. Analysis of the electronic structure shows that the metallacyclic moiety has five occupied π orbitals, two with b1 symmetry and three with a2 symmetry, which describe the π-bonding interactions. The metallabenzenes are thus 10 π-electron systems. This holds for 16-electron and for 18-electron complexes. The π bonding in the metallabenzenes results mainly from the b1 contribution, but the a2 contribution is not negligible. Comparison of the π-bonding strength in the metallacyclic compounds with acylic reference molecules indicates that metallabenzenes should be considered as aromatic compounds whose extra stabilization due to aromatic conjugation is weaker than in benzene. The calculated aromatic stabilization energies (ASEs) are between 8.7 kcal mol−1 for 13 and 37.6 kcal mol−1 for 16 which is nearly as aromatic as benzene (ASE=42.5 kcal mol−1). The classical metallabenzene model compounds 1 and 4 exhibit intermediate aromaticity with ASE values of 33.4 and 17.6 kcal mol−1. The greater stability of the 5d complexes compared with the 4d species appears not to be related to the strength of π conjugation. From the data reported here there is no apparent trend or pattern which indicates a correlation between aromatic stabilization and particular ligands, metals, coordination numbers or charge. The lower metal–C5H5 binding energy of the 4d complexes correlates rather with weaker σ-orbital interactions.
Co-reporter:Ralf Tonner;Greta Heydenrych Dr. Dr.
Chemistry – An Asian Journal 2007 Volume 2(Issue 12) pp:1555-1567
Publication Date(Web):15 OCT 2007
DOI:10.1002/asia.200700235
DFT calculations at the BP86/TZ2P level were carried out to analyze quantitatively the metal–ligand bonding in transition-metal complexes that contain imidazole (IMID), imidazol-2-ylidene (nNHC), or imidazol-4-ylidene (aNHC). The calculated complexes are [Cl4TM(L)] (TM=Ti, Zr, Hf), [(CO)5TM(L)] (TM=Cr, Mo, W), [(CO)4TM(L)] (TM=Fe, Ru, Os), and [ClTM(L)] (TM=Cu, Ag, Au). The relative energies of the free ligands increase in the order IMID<nNHC<aNHC. The energy levels of the carbon σ lone-pair orbitals suggest the trend aNHC>nNHC>IMID for the donor strength, which is in agreement with the progression of the metal–ligand bond-dissociation energy (BDE) for the three ligands for all metals of Groups 4, 6, 8, and 10. The electrostatic attraction can also be decisive in determining trends in ligand–metal bond strength. The comparison of the results of energy decomposition analysis for the Group 6 complexes [(CO)5TM(L)] (L=nNHC, aNHC, IMID) with phosphine complexes (L=PMe3 and PCl3) shows that the phosphine ligands are weaker σ donors and better π acceptors than the NHC tautomers nNHC, aNHC, and IMID.
Co-reporter:Israel Fernández and Gernot Frenking
Chemical Communications 2006 (Issue 48) pp:5030-5032
Publication Date(Web):17 Nov 2006
DOI:10.1039/B613671K
π-Conjugation in several donor-substituted cyanoethynylethenes was estimated using energy decomposition analysis (EDA); very good linear correlations between the ΔEπ values and experimental data are found.
Co-reporter:Helgard G. Raubenheimer, Matthias W. Esterhuysen, Gernot Frenking, Alexey Y. Timoshkin, Catharine Esterhuysen and Ulrike E. I. Horvath
Dalton Transactions 2006 (Issue 38) pp:4580-4589
Publication Date(Web):21 Jul 2006
DOI:10.1039/B607613K
Deprotonated Fischer-type aminocarbene complexes, (CO)5MC(NR2)CH3 (M = Cr or W; R = Me or propyl), react with Ph3PAu+ by metal group substitution – (CO)5M for Ph3PAu+
– and attachment of the extricated M(CO)5 to the deprotonated methyl group. (The products may also be seen as aminovinylgold compounds coordinated to M(CO)5 moieties.) DFT calculations at the B3LYP level of theory using model compounds indicate a clear preference of the gold unit for central C to terminal coordination in the ligand [NMe2CCH2]−, whereas the Cr(CO)5 has a 7 kcal mol−1 preference for C(vinyl) coordination compared to N-coordination. In related thiocarbenes, the sulfur donor atom should be the preferred point of attachment for the metal carbonyl unit. The latter prediction is borne out in practice, and in the three products isolated, including Ph3PAu{C(CH2)SPh}Cr(CO)5 in a mixed crystal with [Ph3PAuSPh]Cr(CO)5, precisely this coordination mode is present. The latter component of the mixed crystal has also been prepared independently of the vinyl one.
Co-reporter:Sabine K. Schneider, Gerrit R. Julius, Christoph Loschen, Helgard G. Raubenheimer, Gernot Frenking and Wolfgang A. Herrmann
Dalton Transactions 2006 (Issue 9) pp:1226-1233
Publication Date(Web):01 Dec 2005
DOI:10.1039/B512419K
A series of cationic pyridinylidene and quinolinylidene complexes of chlorobis(triphenylphosphine)nickel(II) were prepared by oxidative substitution of Ni(PPh3)4 with methylated chloropyridines or chloroquinolines. NMR as well as X-ray crystallographic studies confirmed the trans arrangement of the two phosphines in the products. Calculations, using suitable model compounds at the BP86/TZVP level, clearly differentiate between a standard imidazolylidene complex and new complexes of the NHC-type on the one hand, and new complexes classified as rNHC-types—with the heteroatom distant from the carbene carbon—on the other. The latter form significantly stronger bonds—mainly of an electrostatic nature—with the metal.
Co-reporter:Shigeru Nagase Dr.;Nozomi Takagi Dr.;Akira Sekiguchi Dr. Dr.;Andreas Krapp Dipl.-Chem.
ChemPhysChem 2006 Volume 7(Issue 4) pp:799-800
Publication Date(Web):4 APR 2006
DOI:10.1002/cphc.200500689
Bond order of a silicon-silicon bond (1): The statement that the recently synthesized compound RSiSiR (1) (RSi[C(SiMe3)3]2CHMe2) has rather a double than a triple bond is challenged. Arguments are given which support the interpretation of the bonding situation in terms of two donor–acceptor bonds which are enhanced by one π bond (see picture).
Co-reporter:Andreas Krapp Dipl.-Chem.;F. Matthias Bickelhaupt Dr.
Chemistry - A European Journal 2006 Volume 12(Issue 36) pp:
Publication Date(Web):6 OCT 2006
DOI:10.1002/chem.200600564
The chemical bonds in the diatomic molecules Li2–F2 and Na2–Cl2 at different bond lengths have been analyzed by the energy decomposition analysis (EDA) method using DFT calculations at the BP86/TZ2P level. The interatomic interactions are discussed in terms of quasiclassical electrostatic interactions ΔEelstat, Pauli repulsion ΔEPauli and attractive orbital interactions ΔEorb. The energy terms are compared with the orbital overlaps at different interatomic distances. The quasiclassical electrostatic interactions between two electrons occupying 1s, 2s, 2p(σ), and 2p(π) orbitals have been calculated and the results are analyzed and discussed. It is shown that the equilibrium distances of the covalent bonds are not determined by the maximum overlap of the σ valence orbitals, which nearly always has its largest value at clearly shorter distances than the equilibrium bond length. The crucial interaction that prevents shorter bonds is not the loss of attractive interactions, but a sharp increase in the Pauli repulsion between electrons in valence orbitals. The attractive interactions of ΔEorb and the repulsive interactions of ΔEPauli are both determined by the orbital overlap. The net effect of the two terms depends on the occupation of the valence orbitals, but the onset of attractive orbital interactions occurs at longer distances than Pauli repulsion, because overlap of occupied orbitals with vacant orbitals starts earlier than overlap between occupied orbitals. The contribution of ΔEelstat in most nonpolar covalent bonds is strongly attractive. This comes from the deviation of quasiclassical electron–electron repulsion and nuclear–electron attraction from Coulomb's law for point charges. The actual strength of ΔEelstat depends on the size and shape of the occupied valence orbitals. The attractive electrostatic contributions in the diatomic molecules Li2–F2 come from the s and p(σ) electrons, while the p(π) electrons do not compensate for nuclear–nuclear repulsion. It is the interplay of the three terms ΔEorb, ΔEPauli, and ΔEelstat that determines the bond energies and equilibrium distances of covalently bonded molecules. Molecules like N2 and O2, which are usually considered as covalently bonded, would not be bonded without the quasiclassical attraction ΔEelstat.
Co-reporter:Beatrice Buchin;Christian Gemel Dr.;Thomas Cadenbach;Israel Fernández Dr. Dr.;Rol A. Fischer Dr.
Angewandte Chemie 2006 Volume 118(Issue 31) pp:
Publication Date(Web):6 JUL 2006
DOI:10.1002/ange.200601007
Mit der Ga+-Quelle [Ga2Cp*][BArF] reagiert [Pt(GaCp*)4] zum Titelkomplex, dessen „nacktes“ Galliumatom terminal an das Platinzentrum koordiniert ist. DFT-Rechungen ergaben, dass Ga+ als σ-Acceptor wirkt und dass die Pt-Ga-Bindung außerdem durch π-Wechselwirkungen verstärkt wird. Cp*=C5Me5; BArF=B{C6H3-3,5-(CF3)2}4.
Co-reporter:Gernot Frenking ;Catherine Esterhuysen Dr.;Attila Kovacs
Chemistry - A European Journal 2006 Volume 12(Issue 30) pp:
Publication Date(Web):12 SEP 2006
DOI:10.1002/chem.200600909
In our reply to the preceding comment by Richard Bader we show that the statements of the author are not justified and that he contradicts his own previous work.
Co-reporter:Beatrice Buchin;Christian Gemel Dr.;Thomas Cadenbach;Israel Fernández Dr. Dr.;Rol A. Fischer Dr.
Angewandte Chemie International Edition 2006 Volume 45(Issue 31) pp:
Publication Date(Web):6 JUL 2006
DOI:10.1002/anie.200601007
In the buff: The reaction of [Pt(GaCp*)4] with the Ga+ source [Ga2Cp*][BArF] (Cp*=C5Me5; BArF=B{C6H3(3,5-CF3)2}4) leads to the title complex, which exhibits a naked gallium atom coordinated to a platinum center. DFT calculations reveal that the Ga+ ligand serves as a σ acceptor, and the PtGa bond is additionally supported by strong π interactions.
Co-reporter:Ralf Tonner;Florian Öxler;Bernhard Neumüller Dr.;Wolfgang Petz Dr. Dr.
Angewandte Chemie International Edition 2006 Volume 45(Issue 47) pp:
Publication Date(Web):31 OCT 2006
DOI:10.1002/anie.200602552
A combined experimental/theoretical study gives strong evidence that carbodiphosphoranes are divalent carbon(0) compounds. The calculations show that carbodiphosphoranes have two lone pairs of electrons (see picture), which give rise to unusual properties as confirmed by experiment. The synthesis of a triply charged molecules in which two protonated carbodiphosphoranes serve as donor ligands to an Ag+ center supports the bonding model.
Co-reporter:Israel Fernández Dr.
Chemistry - A European Journal 2006 Volume 12(Issue 13) pp:
Publication Date(Web):27 FEB 2006
DOI:10.1002/chem.200501405
The intrinsic strength of π interactions in conjugated and hyperconjugated molecules has been calculated using density functional theory by energy decomposition analysis (EDA) of the interaction energy between the conjugating fragments. The results of the EDA of the trans-polyenes H2CCH(HCCH)nCHCH2 (n=1–3) show that the strength of π conjugation for each CC moiety is higher than in trans-1,3-butadiene. The absolute values for the conjugation between SiSi π bonds are around two-thirds of the conjugation between CC bonds but the relative contributions of ΔEπ to ΔEorb in the all-silicon systems are higher than in the carbon compounds. The π conjugation between CC and CO or CNH bonds in H2CCHC(H)O and H2CCHC(H)NH is comparable to the strength of the conjugation between CC bonds. The π conjugation in H2CCHC(R)O decreases when R=Me, OH, and NH2 while it increases when R=halogen. The hyperconjugation in ethane is around a quarter as strong as the π conjugation in ethyne. Very strong hyperconjugation is found in the central CC bonds in cubylcubane and tetrahedranyltetrahedrane. The hyperconjugation in substituted ethanes X3CCY3 (X,Y=Me, SiH3, F, Cl) is stronger than in the parent compound particularly when X,Y=SiH3 and Cl. The hyperconjugation in donor–acceptor-substituted ethanes may be very strong; the largest ΔEπ value was calculated for (SiH3)3CCCl3 in which the hyperconjugation is stronger than the conjugation in ethene. The breakdown of the hyperconjugation in X3CCY3 shows that donation of the donor-substituted moiety to the acceptor group is as expected the most important contribution but the reverse interaction is not negligible. The relative strengths of the π interactions between two CC double bonds, one CC double bond and CH3 or CMe3 substituents, and between two CH3 or CMe3 groups, which are separated by one CC single bond, are in a ratio of 4:2:1. Very strong hyperconjugation is found in HCCC(SiH3)3 and HCCCCl3. The extra stabilization of alkenes and alkynes with central multiple bonds over their terminal isomers coming from hyperconjugation is bigger than the total energy difference between the isomeric species. The hyperconjugation in MeC(R)O is half as strong as the conjugation in H2CCHC(R)O and shows the same trend for different substituents R. Bond energies and lengths should not be used as indicators of the strength of hyperconjugation because the effect of σ interactions and electrostatic forces may compensate for the hyperconjugative effect.
Co-reporter:Stefan Erhardt Dr.
Chemistry - A European Journal 2006 Volume 12(Issue 17) pp:
Publication Date(Web):6 APR 2006
DOI:10.1002/chem.200500580
The equilibrium geometries and bond energies of the complexes H3BL and H2B+L (L=CO; EC5H5: E=N, P, As, Sb, Bi) have been calculated at the BP86/TZ2P level of theory. The nature of the donor–acceptor bonds was investigated by energy decomposition analysis (EDA). The bond strengths of H3BL have the order CO>N>P>As>Sb>Bi. The calculated values are between De=37.1 kcal mol−1 for H3BCO and De=6.9 kcal mol−1 for H3BBiC5H5. The bond dissociation energies of the cations H2B+CO and H2B+EC5H5 are larger than for H3BL, particularly for complexes of the heterobenzene ligands. The calculated values are between De=51.9 kcal mol−1 for H2B+CO and De=122.1 kcal mol−1 for H2B+NC5H5. The trend of the BDE of H2B+CO and H2B+EC5H5 is N>P>As>Sb>Bi>CO. A surprising result is found for H2B+CO, which has a significantly stronger and yet substantially longer bond than H3BCO. The reason for the longer but stronger bond in H2B+CO compared with that in H3BCO comes mainly from the change in electrostatic attraction and π bonding at shorter distances, which increases more in the neutral system than in the cation, and to a lesser extent from the deformation energy of the fragments. The H2B+NC5H5 π⊥ donation plays an important role for the stronger interactions at shorter distances compared with those in H3BNC5H5. The attractive interaction in H2B+CO further increases at bond lengths that are shorter than the equilibrium value, but this is compensated by the energy which is necessary to deform BH2+ from its linear equilibrium geometry to the bent form in the complex. The EDA shows that the contributions of the orbital interactions to the donor–acceptor bonds are always larger than the classical electrostatic contributions, but the latter term plays an important role for the trend in bond strength. The largest contributions to the orbital interactions come from the σ orbitals. The EDA calculations suggest that heterobenzene ligands may become moderately strong π donors in complexes with strong Lewis acids, while CO is only a weak π donor. The much stronger interaction energies in H2B+EC5H5 compared with those in H3BEC5H5 are caused by the significantly larger contribution of the π⊥ orbitals in H2B+EC5H5 and by the increase of the binding interactions of the σ+π∥ orbitals.
Co-reporter:Sabine K. Schneider;Patric Roembke;Gerrit R. Julius;Christoph Loschen;Helgard G. Raubenheimer;Wolfgang A. Herrmann
European Journal of Inorganic Chemistry 2005 Volume 2005(Issue 15) pp:
Publication Date(Web):28 JUN 2005
DOI:10.1002/ejic.200500307
A palladium complex, trans-Cl(PPh3)2Pd{Ca-cyclo-C6H4-o-N(Me)C=CbH}(Ca—Cb), that contains a carbene ligand with remote heteroatoms (rNHC) is much more active in certain C–C coupling reactions than comparably simple NHC- and phosphane-containing precatalysts. The rNHC ligand binds strongly by electrostatic as well as orbital interaction to the metal. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2005)
Co-reporter:Gernot Frenking, Miquel Solà, Sergei F. Vyboishchikov
Journal of Organometallic Chemistry 2005 Volume 690(24–25) pp:6178-6204
Publication Date(Web):1 December 2005
DOI:10.1016/j.jorganchem.2005.08.054
In this work, we summarize recent theoretical studies of our groups in which modern quantum chemical methods are used to gain insight into the nature of metal–ligand interactions in Fischer- and Schrock-type carbene complexes. It is shown that with the help of charge- and energy-partitioning techniques it is possible to build a bridge between heuristic bonding models and the physical mechanism which leads to a chemical bond. Questions about the bonding situation which in the past were often controversially discussed because of vaguely defined concepts may be addressed in terms of well defined quantum chemical expressions. The results of the partitioning analyses show that Fischer and Schrock carbenes exhibit different bonding situations, which are clearly revealed by the calculated data. The contributions of the electrostatic and the orbital interaction are estimated and the strength of the σ donor and π acceptor bonding in Fischer complexes are discussed. We also discuss the bonding situation in complexes with N,N-heterocyclic carbene ligands.The nature of the chemical bonding in Fischer- and Schrock-type carbene complexes as well as in complexes with N,N-heterocyclic carbene ligands has been analyzed with charge- and energy-partitioning methods.
Co-reporter:Daniel Cappel;Sor Tüllmann;Andreas Krapp Dipl.-Chem. Dr.
Angewandte Chemie 2005 Volume 117(Issue 23) pp:
Publication Date(Web):4 MAY 2005
DOI:10.1002/ange.200500452
Co-reporter:Markus Hölscher Dr.;Walter Leitner ;Max C. Holthausen Dr.
Chemistry - A European Journal 2005 Volume 11(Issue 16) pp:
Publication Date(Web):25 MAY 2005
DOI:10.1002/chem.200500217
Quantum chemical calculations by using density functional theory at the B3LYP level have been carried out to elucidate the reaction course for the addition of ethylene to [OsO2(CH2)2] (1). The calculations predict that the kinetically most favorable reaction proceeds with an activation barrier of 8.1 kcal mol−1 via [3+2] addition across the OOsCH2 moiety. This reaction is −42.4 kcal mol−1 exothermic. Alternatively, the [3+2] addition to the H2COsCH2 fragment of 1 leads to the most stable addition product 4 (−72.7 kcal mol−1), yet this process has a higher activation barrier (13.0 kcal mol−1). The [3+2] addition to the OOsO fragment yielding 2 is kinetically (27.5 kcal mol−1) and thermodynamically (−7.0 kcal mol−1) the least favorable [3+2] reaction. The formal [2+2] addition to the OsO and OsCH2 double bonds proceeds by initial rearrangement of 1 to the metallaoxirane 1 a. The rearrangement 11 a and the following [2+2] additions have significantly higher activation barriers (>30 kcal mol−1) than the [3+2] reactions. Another isomer of 1 is the dioxoosmacyclopropane 1 b, which is 56.2 kcal mol−1 lower in energy than 1. The activation barrier for the 11 b isomerization is 15.7 kcal mol−1. The calculations predict that there are no energetically favorable addition reactions of ethylene with 1 b. The isomeric form 1 c containing a peroxo group is too high in energy to be relevant for the reaction course. The accuracy of the B3LYP results is corroborated by high level post-HF CCSD(T) calculations for a subset of species.
Co-reporter:Attila Kovács Dr.;Catharine Esterhuysen Dr.
Chemistry - A European Journal 2005 Volume 11(Issue 6) pp:
Publication Date(Web):25 JAN 2005
DOI:10.1002/chem.200400525
The nature of the chemical bond in nonpolar molecules has been investigated by energy-partitioning analysis (EPA) of the ADF program using DFT calculations. The EPA divides the bonding interactions into three major components, that is, the repulsive Pauli term, quasiclassical electrostatic interactions, and orbital interactions. The electrostatic and orbital terms are used to define the nature of the chemical bond. It is shown that nonpolar bonds between main-group elements of the first and higher octal rows of the periodic system, which are prototypical covalent bonds, have large attractive contributions from classical electrostatic interactions, which may even be stronger than the attractive orbital interactions. Fragments of molecules with totally symmetrical electron-density distributions, like the nitrogen atoms in N2, may strongly attract each other through classical electrostatic forces, which constitute 30.0 % of the total attractive interactions. The electrostatic attraction can be enhanced by anisotropic charge distribution of the valence electrons of the atoms that have local areas of (negative) charge concentration. It is shown that the use of atomic partial charges in the analysis of the nature of the interatomic interactions may be misleading because they do not reveal the topography of the electronic charge distribution. Besides dinitrogen, four groups of molecules have been studied. The attractive binding interactions in HnEEHn (E=Li to F; n=0–3) have between 20.7 (E=F) and 58.4 % (E=Be) electrostatic character. The substitution of hydrogen by fluorine does not lead to significant changes in the nature of the binding interactions in FnEEFn (E=Be to O). The electrostatic contributions to the attractive interactions in FnEEFn are between 29.8 (E=O) and 55.3 % (E=Be). The fluorine substituents have a significant effect on the Pauli repulsion in the nitrogen and oxygen compounds. This explains why F2NNF2 has a much weaker bond than H2NNH2, whereas the interaction energy in FOOF is much stronger than in HOOH. The orbital interactions make larger contributions to the double bonds in HBBH, H2CCH2, and HNNH (between 59.9 % in B2H2 and 65.4 % in N2H2) than to the corresponding single bonds in HnEEHn. The orbital term ΔEorb (72.4 %) makes an even greater contribution to the HCCH triple bond. The contribution of ΔEorb to the HnEEHn bond increases and the relative contribution of the π bonding decreases as E becomes more electronegative. The π-bonding interactions in HCCH amount to 44.4 % of the total orbital interactions. The interaction energy in H3EEH3 (E=C to Pb) decreases monotonically as the element E becomes heavier. The electrostatic contributions to the EE bond increases from E=C (41.4 %) to E=Sn (55.1 %) but then decreases when E=Pb (51.7 %). A true understanding of the strength and trends of the chemical bonds can only be achieved when the Pauli repulsion is considered. In an absolute sense the repulsive ΔEPauli term is in most cases the largest term in the EPA.
Co-reporter:Daniel Cappel;Sor Tüllmann;Andreas Krapp Dipl.-Chem. Dr.
Angewandte Chemie International Edition 2005 Volume 44(Issue 23) pp:
Publication Date(Web):4 MAY 2005
DOI:10.1002/anie.200500452
Co-reporter:Stefan Erhardt Dipl.-Chem. Dr.;Zhongfang Chen Dr.;Paul von Ragué Schleyer
Angewandte Chemie International Edition 2005 Volume 44(Issue 7) pp:
Publication Date(Web):21 JAN 2005
DOI:10.1002/anie.200461970
Computational exploration of carboranes of formula CnBmq+ reveals that more than one hypercoordinated carbon atom can be enclosed by a peripheral ring comprising a suitable number of boron atoms. The C2B8, C3B93+, and C5B11+ species (the LUMO of the latter is shown) are stabilized by substantial Hückel π aromaticity. Furthermore, multicenter σ bonding helps bind the inner carbon units to the boron perimeters, though they can freely rotate relative to one another.
Co-reporter:Jingping Zhang, Gernot Frenking
Chemical Physics Letters 2004 Volume 394(1–3) pp:120-125
Publication Date(Web):11 August 2004
DOI:10.1016/j.cplett.2004.06.074
Abstract
Geometries of ground and first excited states of mer-tris(8-hydroxyquinolinato)metal (Mq3, M = AlIII, GaIII) are optimized at B3LYP/6-31G(d) and CIS/6-31G level, respectively. In order to investigate the difference for individual ligands in mer-Mq3, the energy partitioning analysis has been carried out at the BP86 level using TZ2P basis functions for the bonding interactions between each fragment Mq2 and single ligand q. HOMO and LUMO distribution fashion can be traced back to the lowest electrostatic attractive and highest orbital interaction energy between fragments A-quinolate ligand and Mq2 and B-ligand and Mq2, respectively.
Co-reporter:Gernot Frenking Dr.
Angewandte Chemie International Edition 2003 Volume 42(Issue 29) pp:
Publication Date(Web):24 JUL 2003
DOI:10.1002/anie.200320070
Co-reporter:Matthias Lein Dipl.-Chem.;Jan Frunzke Dipl.-Chem. Dr.
Angewandte Chemie International Edition 2003 Volume 42(Issue 11) pp:
Publication Date(Web):13 MAR 2003
DOI:10.1002/anie.200390336
Iron enters the pentagon: Quantum chemical calculations using gradient-corrected DFT predict that the cations [Fe(Sb5)]+ and [Fe(Bi5)]+ in the electronic singlet state have planar (D5h) equilibrium geometries (see scheme). Analysis of the electronic structure shows that the molecules are metal-centered six-π-electron aromatic species with strong iron–ligand π bonds which involve the d(π) atomic orbitals of the Fe center and the degenerate π orbital of the ring. The calculated 57Fe NMR chemical shifts indicate extremely high deshielding of the metal nucleus.
Co-reporter:Catharine Esterhuysen Dr.
Chemistry - A European Journal 2003 Volume 9(Issue 15) pp:
Publication Date(Web):28 JUL 2003
DOI:10.1002/chem.200304723
Complexes of W(CO)5 with neutral diatomic pnictogen ligands N2, P2, As2, Sb2, and Bi2 and anionic Group 14 ligands Si22−, Ge22−, Sn22−, and Pb22− coordinated in both side-on and end-on fashion have been optimized by using density functional theory at the BP86 level with valence sets of TZP quality. The calculated bond energies have been used to compare the preferential binding modes of each respective ligand. The results were interpreted by analyzing the nature of the interaction between the ligands and the metal fragment using an energy partitioning method. This yields quantitative information regarding the strength of covalent and electrostatic interactions between the metal and ligand, as well as the contributions by orbitals of different symmetry to the covalent bonding. Results show that all the ligands studied bind preferentially in a side-on coordination mode, with the exception of N2, which prefers to coordinate in an end-on mode. The preference of the heavier homologues P2–Bi2 for binding in a side-on mode over the end-on mode in the neutral complexes [(CO)5WE2] comes mainly from the much stronger electrostatic attraction in the former species. The energy difference between the side-on and end-on isomers of the negatively charged complexes with the ligands Si22−, Ge22−, Sn22−, and Pb22− is much less and it cannot be ascribed to a particular bonding component.
Co-reporter:Gernot Frenking Dr.
Angewandte Chemie 2003 Volume 115(Issue 29) pp:
Publication Date(Web):24 JUL 2003
DOI:10.1002/ange.200320070
Co-reporter:Víctor M. Rayón Dr.
Chemistry - A European Journal 2002 Volume 8(Issue 20) pp:
Publication Date(Web):14 OCT 2002
DOI:10.1002/1521-3765(20021018)8:20<4693::AID-CHEM4693>3.0.CO;2-B
The geometries, metal–ligand bond dissociation energies, and heats of formation of twenty sandwich and half-sandwich complexes of the main-group elements of Groups 1, 2, 13, and 14, and Zn have been calculated with quantum chemical methods. The geometries of the [E(Cp)] and [E(Cp)2] complexes were optimized using density functional theory at the BP86 level with valence basis sets, which have DZP and TZP quality. Improved energy values have been obtained by using coupled-cluster theory at the CCSD(T) level. The nature of the metal–ligand bonding has been analyzed with an energy-partitioning method. The results give quantitative information about the strength of the covalent and electrostatic interactions between En+ and (Cp−)n (n=1, 2). The contributions of the orbitals with different symmetry to the covalent bonding are also given.
Co-reporter:Yu Chen;Michael Hartmann
European Journal of Inorganic Chemistry 2001 Volume 2001(Issue 6) pp:
Publication Date(Web):23 APR 2001
DOI:10.1002/1099-0682(200106)2001:6<1441::AID-EJIC1441>3.0.CO;2-G
Density functional theory and high-level ab initio calculations are used to evaluate the influence of mono- and dinuclear iron carbonyl complexes to the fixation and stepwise hydrogenation of dinitrogen via diazene and hydrazine to ammonia. In comparison to the reaction of isolated N2, only the first step in this reaction sequence (i.e. the reduction of N2 to N2H2) experiences a significant change in its thermochemistry when coordinated to mono- or dinuclear iron tetracarbonyl fragments. The reaction enthalpy ΔHR0 (T = 0 K) for the endothermic hydrogenation of (CO)4Fe−N2 to give (CO)4Fe−N2H2 is lower than for the corresponding metal-free process by 16.1 kcal mol−1. The analogous step involving the dinuclear species (CO)4Fe−N2−Fe(CO4) and (CO)4Fe−N2H2−Fe(CO)4 is even less endothermic than the reduction involving only one iron tetracarbonyl complex by 13.1 kcal mol−1. In comparison to that, the second and third step of this reduction sequence, namely the conversion of coordinated diazene to (CO)4Fe−N2H4 and the subsequent reduction of coordinated hydrazine to (CO)4Fe−NH3 show only relatively small thermodynamic changes compared to the analogous reactions of isolated N2H2 and N2H4. The reduction of (CO)4Fe−N2H2 to (CO)4Fe−N2H4 is almost as exothermic as the analogue reaction involving isolated N2H2, whereas the hydrogenation of (CO)4Fe−N2H4 to (CO)4Fe−NH3 is less exothermic by 4.0 kcal mol−1. Finally, the reduction of (CO)4Fe−N2H2−Fe(CO)4 and (CO)4Fe−N2H4−Fe(CO)4 are both predicted to be less exothermic than their mononuclear analogues by 4.0 and 1.1 kcal mol−1, respectively. Moreover, we find that only N2 and N2H2, which already show a noticeable π-acceptor behavior in their complexes with Fe(CO)4, experience important structural changes in their corresponding dinuclear complexes, i.e. a shortening of the Fe−N bonds and a lengthening of the N−N bonds on going from (CO)4Fe−L to (CO)4Fe−L−Fe(CO)4 (L = N2, N2H2). This behavior is in line with a slightly increased π-acceptor ability of these ligands in their respective dinuclear complexes. Such structural changes are absent for N2H4, which only exhibits a comparatively weak π-acceptor character in (CO)4Fe−N2H4.
Co-reporter:Gernot Frenking
Journal of Organometallic Chemistry 2001 Volume 635(1–2) pp:9-23
Publication Date(Web):15 October 2001
DOI:10.1016/S0022-328X(01)01154-8
The results of an energy partitioning analysis of three classes of transition metal complexes are discussed. They are (i) neutral and charged isoelectronic hexacarbonyls TM(CO)6q (TMq=Hf2−, Ta−, W, Re+, Os2+, Ir3+); (ii) Group-13 diyl complexes (CO)4FeER (E=B, Al, Ga, In, Tl; R=Cp, Ph), Fe(ECH3)5 and Ni(ECH3)4; (iii) complexes with cyclic π-donor ligands Fe(Cp)2 and Fe(η5-N5)2. The results show that Dewar's molecular orbital model can be recovered and that the orbital interactions can become quantitatively expressed by accurate quantum chemical calculations. However, the energy analysis goes beyond the MO model and gives a much deeper insight into the nature of the metal–ligand bonding. It addresses also the question of ionic versus covalent bonding as well as the relative importance of σ and π bonding contributions.The bonding model suggested for metal–olefin complexes, which was suggested by Dewar 50 years ago, is found by an energy partitioning analysis to be a valid description of the bonding in transition metal complexes with ligands CO, Group-13 diyl species ER (E=B, Al, Ga, In, Tl; R=Cp, Ph, CH3) and with cyclic π-donor ligands Cp and cyc-N5 in Fe(Cp)2 and Fe (η5-N5)2.
Co-reporter:Nicole Dölker, Gernot Frenking
Journal of Organometallic Chemistry 2001 Volumes 617–618() pp:225-232
Publication Date(Web):15 January 2001
DOI:10.1016/S0022-328X(00)00554-4
[RuH2Cl2(PiPr3)2] reacts with terminal alkynes to give the vinylidene complex [RuCl2(CCHR)(PiPr3)2]. As a side product the carbene complex [RuCl2(CHR)(PiPr3)2] is formed. The formation of the vinylidene compound has been studied widely and is well understood whereas the reaction mechanism that leads to the carbene complex is still unclear. We have studied two possible reaction paths at the B3LYP level of theory: on the one hand the addition of acetylene and two subsequent 1,3-H shifts from the metal center to the C2 carbon of the acetylene ligand; on the other hand the dissociation of HCl from the starting compound, rearrangement of acetylene to vinyl and the formation of the carbene by addition of HCl. Both reaction paths have been found to be possible. The former can be understood as a 1,3-H shift followed by a 1,2-H shift due to the unusual η2 coordination mode of the vinyl intermediate. The latter proceeds via protonation of the vinyl ligand and addition of Cl− to the metal center.
Co-reporter:Yu Chen M. Sc.;Michael Hartmann Dr.;Michael Diedenhofen Dr. Dr.
Angewandte Chemie 2001 Volume 113(Issue 11) pp:
Publication Date(Web):28 MAY 2001
DOI:10.1002/1521-3757(20010601)113:11<2107::AID-ANGE2107>3.0.CO;2-U
Quantenchemische Rechnungen auf dem B3LYP-Niveau zeigen, dass die beiden sterisch anspruchsvollen Substituenten Ar*=2,6-Ph2C6H3 in Ar*PbPbAr* die gezeigte trans-gewinkelte Gleichgewichtsstruktur als niedrigst liegendes Energieminimum erzwingen. In Ar*PbPbAr* liegt keine Pb-Pb-π-Bindung vor. Bei den Energieminima der Stammverbindungen PhPbPbPh und HPbPbH handelt es sich dagegen um zweifach verbrückte Strukturen.
Co-reporter:Matthias Lein C. Chem.;Jan Frunzke Dipl. Chem.;Alexey Timoshkin Dr.
Chemistry - A European Journal 2001 Volume 7(Issue 19) pp:
Publication Date(Web):27 SEP 2001
DOI:10.1002/1521-3765(20011001)7:19<4155::AID-CHEM4155>3.0.CO;2-M
Quantum-chemical calculations with gradient-corrected (B3LYP) density functional theory have been carried out for iron bispentazole and ferrocene. The calculations predict that Fe(η5- N5)2 is a strongly bonded complex which has D5d symmetry. The theoretically predicted total bond energy that yields Fe in the 5D ground state and two pentazole ligands is Do=109.0 kcal mol−1, which is only 29 kcal mol−1 less than the calculated bond energy of ferrocene (Do=138.0 kcal mol−1; experimental: 158±2 kcal mol−1). The compound Fe(η5-N5)2 is 260.5 kcal mol−1 higher in energy than the experimentally known isomer Fe(N2)5, but the bond energy of the latter (Do=33.7 kcal mol−1) is much less. The energy decomposition analyses of Fe(η5-N5)2 and ferrocene show that the two compounds have similar bonding situations. The metal–ligand bonds are roughly half ionic and half covalent. The covalent bonding comes mainly from (e1g) η5-N5−Fe2+ π-donation. The previously suggested MO correlation diagram for ferrocene is nicely recovered by the Kohn–Sham orbitals. The calculated vibrational frequencies and IR intensities are reported.
Co-reporter:Jörg Rissler Dipl.-Chem.;Michael Hartmann Dr.;Christina M. March Dr.;Hansjörg Grützmacher Dr. Dr.
Chemistry - A European Journal 2001 Volume 7(Issue 13) pp:
Publication Date(Web):27 JUN 2001
DOI:10.1002/1521-3765(20010702)7:13<2834::AID-CHEM2834>3.0.CO;2-E
Quantum chemical calculations at the MP2 and CCSD(T) levels of theory are reported for cations of the general type [A(XH2)3]+ with A=C, Si and X=N, P, As, Sb, Bi. Population analysis, methyl stabilization energies (MSEs), and structural criteria were used to predict the p(π)-donor ability of and the π-stabilization energy exerted by this series of pnicogens. All of the substituents XH2 considered in these studies invariably stabilize the triply substituted carbenium as well as the silicenium ions. The calculated data show that the intrinsic p(π)-donation of the group 15 atoms follows the order N<P<As<Sb<Bi. However, the trend of the stabilization energies is fully reversed. The intrinsic stabilization energies of the planar carbenium ions decrease monotonically from 161.2 kcal mol−1 for X=NH2 to 98.0 kcal mol−1 for X=BiH2. The effective stabilization of the pnicogens in the equilibrium structures, which also includes the energy-demanding pyramidalization of the XH2 substituents, follows the same trend, although the absolute numbers are reduced to 145.6 kcal mol−1 for X=NH2 and 53.2 kcal mol−1 for X=BiH2. This seemingly contrasting behavior of increasing p(π) charge donation and decreasing stabilization has already been found for other substituents. Previous studies have shown that carbenium ions substituted by chalcogens up to the fourth row also stabilize C+ less effectively with respect to heavier substituents. Of the ions investigated in this study, only the silicenium ions that are stabilized by pnicogens from the third to the sixth row of the periodic system yield increased stabilizing energies that follow the corresponding intrinsic p(π)-donor abilities of the respective substituent.
Co-reporter:Dana Weiß, Manuela Winter, Roland A. Fischer, Chen Yu, Karin Wichmann and Gernot Frenking
Chemical Communications 2000 (Issue 24) pp:2495-2496
Publication Date(Web):30 Nov 2000
DOI:10.1039/B008133G
A new homoleptic diplatinum complex
[Pt2(GaCp*)2(μ2-GaCp*)3]
(Cp* = pentamethylcyclopentadienyl) exhibiting a central unit of two
platinum atoms coordinated by five Cp*Ga groups acting as terminal as well
as bridging ligands, was synthesized by the reaction of
tris(ethylene)platinum(0) with an excess of (Cp*Ga)6
and was characterized by structural and quantum chemical methods.
Co-reporter:Ariana Beste;Oliver Krämer;Anja Gerhard
European Journal of Inorganic Chemistry 1999 Volume 1999(Issue 11) pp:
Publication Date(Web):20 OCT 1999
DOI:10.1002/(SICI)1099-0682(199911)1999:11<2037::AID-EJIC2037>3.0.CO;2-T
Quantum chemical calculations at the MP2 level using large valence basis sets up to TZ+2P quality have been carried out in order to predict the geometries and bond energies of the title compounds. The nature of the donor–acceptor bond has also been investigated. The calculations show clearly that diaminocarbenes are much stronger Lewis bases than amines. The complexation energies of C(NH2)2 have been calculated to be 14–27 kcal/mol higher than those of NH3. The most strongly bonded complex is Cl3Al–C(NH2)2, which has a theoretically predicted Al–C bond energy Do = 59.1 kcal/mol. In all the complexes, the strength of the Lewis bases is C(NH2)2 > NH3 > CO, but the ordering of Lewis acid strength of EX3 depends on the coordinated Lewis base. TiF4 and TiCl4 have similar Lewis acidities as BF3, but the titanium tetrahalides may bind one or two donor molecules with almost the same bond strength. The investigated donor–acceptor bonds have a high degree of ionic character. The largest covalent contributions are found for the diaminocarbene complexes. The covalent character of the X3E–CO bond increases on going from E = boron to the heavier Group 13 elements, while the opposite order is found for the X3E–NH3 and X3E–C(NH2)2 bonds.
Co-reporter:Anthony J. Lupinetti;Volker Jonas;Walter Thiel;Steven H. Strauss
Chemistry - A European Journal 1999 Volume 5(Issue 9) pp:
Publication Date(Web):30 AUG 1999
DOI:10.1002/(SICI)1521-3765(19990903)5:9<2573::AID-CHEM2573>3.0.CO;2-J
Surprising trendsin the M–CO bond dissociation energies (see illustration; n = number of carbonyl ligands) are revealed by quantum chemical calculations for the title molecules. Analysis of the bonding interactions reveals an interplay of coulombic and covalent interactions between the metal cations and CO. Theoretical predictions are made for strongly bound carbonyl complexes which have not yet been synthesized.
Co-reporter:Christian Boehme
Chemistry - A European Journal 1999 Volume 5(Issue 7) pp:
Publication Date(Web):24 JUN 1999
DOI:10.1002/(SICI)1521-3765(19990702)5:7<2184::AID-CHEM2184>3.0.CO;2-3
Gallium–iron bonds in the model compounds [(C6H5)GaFe(CO)4] (1 a) and [CpGaFe(CO)4] (2 a) were analyzed using the CDA partitioning scheme. The Ga−Fe bonds are largely ionic. There is a substantially higher degree of GaFe π backbonding in 1 a than in 2 a. Gallium becomes stabilized in 1 a mainly by GaFe backdonation, whereas GaCp backdonation stabilizes Ga in 2 a.
Co-reporter:Gernot Frenking, Ralf Tonner, Susanne Klein, Nozomi Takagi, Takayazu Shimizu, Andreas Krapp, Krishna K. Pandey and Pattiyil Parameswaran
Chemical Society Reviews 2014 - vol. 43(Issue 14) pp:NaN5139-5139
Publication Date(Web):2014/06/11
DOI:10.1039/C4CS00073K
Recent theoretical studies are reviewed which show that the naked group 14 atoms E = C–Pb in the singlet 1D state behave as bidentate Lewis acids that strongly bind two σ donor ligands L in the donor–acceptor complexes L→E←L. Tetrylones EL2 are divalent E(0) compounds which possess two lone pairs at E. The unique electronic structure of tetrylones (carbones, silylones, germylones, stannylones, plumbylones) clearly distinguishes them from tetrylenes ER2 (carbenes, silylenes, germylenes, stannylenes, plumbylenes) which have electron-sharing bonds R–E–R and only one lone pair at atom E. The different electronic structures of tetrylones and tetrylenes are revealed by charge- and energy decomposition analyses and they become obvious experimentally by a distinctively different chemical reactivity. The unusual structures and chemical behaviour of tetrylones EL2 can be understood in terms of the donor–acceptor interactions L→E←L. Tetrylones are potential donor ligands in main group compounds and transition metal complexes which are experimentally not yet known. The review also introduces theoretical studies of transition metal complexes [TM]–E which carry naked tetrele atoms E = C–Sn as ligands. The bonding analyses suggest that the group-14 atoms bind in the 3P reference state to the transition metal in a combination of σ and π∥ electron-sharing bonds TM–E and π⊥ backdonation TM→E. The unique bonding situation of the tetrele complexes [TM]–E makes them suitable ligands in adducts with Lewis acids. Theoretical studies of [TM]–E→W(CO)5 predict that such species may becomes synthesized.
Co-reporter:Milind M. Deshmukh, Shridhar R. Gadre, Ralf Tonner and Gernot Frenking
Physical Chemistry Chemical Physics 2008 - vol. 10(Issue 17) pp:NaN2301-2301
Publication Date(Web):2008/03/10
DOI:10.1039/B803068E
The molecular electrostatic potentials of divalent carbon(0) and divalent carbon(II) compounds are calculated and the results are compared with theoretically predicted proton affinities and complexation energies with BH3.
Co-reporter:Gernot Frenking, Markus Hermann, Diego M. Andrada and Nicole Holzmann
Chemical Society Reviews 2016 - vol. 45(Issue 4) pp:NaN1144-1144
Publication Date(Web):2016/01/27
DOI:10.1039/C5CS00815H
A summary of theoretical and experimental work in the area of low-coordinated compounds of boron and group-14 atoms C–Sn in the last decade is presented. The focus of the account lies on molecules EL2, E2L2 and E3L3, which possess dative bonds between one, two or three atoms E and σ-donor ligands L that stabilize the atoms E through L→E donor–acceptor interactions. The interplay between theory and experiment provides detailed insight into the bonding situation of the molecules, which serves as guideline for the synthesis of molecules that possess unusual bonding motifs.
Co-reporter:Ulrich Siemeling, Christian Färber, Clemens Bruhn, Michael Leibold, Detlef Selent, Wolfgang Baumann, Moritz von Hopffgarten, Catharina Goedecke and Gernot Frenking
Chemical Science (2010-Present) 2010 - vol. 1(Issue 6) pp:NaN704-704
Publication Date(Web):2010/10/22
DOI:10.1039/C0SC00451K
N-Heterocyclic carbenes (NHCs) are extremely valuable as nucleophilic organocatalysts. They are widely applied as ligands in transition-metal catalysed reactions, where they are known as particularly potent σ-donors. They are commonly viewed as workhorses exhibiting reliable, but undramatic, chemical behaviour. The N → Ccarbene π-donation stabilises NHCs at the expense of low reactivity towards nucleophiles. In contrast to NHCs, stable (alkyl)(amino)carbenes exhibit spectacular reactivity, allowing, for example, the splitting of hydrogen and ammonia and the fixation of carbon monoxide. NHCs have been judged to be electronically not suitable for showing similar reactivity. Here, we demonstrate that a ferrocene-based NHC is able to add ammonia, methyl acrylate, tert-butyl isocyanide, and carbon monoxide—reactions typical of (alkyl)(amino)carbenes, but unprecedented for diaminocarbenes. We also show that even the simplest stable diaminocarbene, C(NiPr2)2, adds CO. This reaction affords a β-lactam by a subsequent intramolecular process involving a C–H activation. Our results shed new light on the chemistry of diaminocarbenes and offer great potential for synthetic chemistry and catalysis.
Co-reporter:Ranajit Saha, Sudip Pan, Gernot Frenking, Pratim K. Chattaraj and Gabriel Merino
Physical Chemistry Chemical Physics 2017 - vol. 19(Issue 3) pp:NaN2293-2293
Publication Date(Web):2016/12/14
DOI:10.1039/C6CP06824C
A coupled-cluster study is performed on CO bound BeY complexes (Y = O, CO3, SO4, NH, NCN, and NBO) to understand the effect of attached ligands (Y) on the CO binding ability and C–O stretching frequency (νCO). Herein, we report that BeNCN has the highest CO binding ability (via both C- and O-side binding) among the studied neutral Be-based clusters, whereas OCBeSO4 has the highest νCO among the neutral carbonyls. The nature and extent of shift in νCO compared to free CO are explained in terms of change in polarization in the bonding orbitals of CO and relative contribution from OC→BeY or CO→BeY σ-donation, and OC←BeY or CO←BeY π-back-donation. The largest blue-shift in OCBeSO4 and the largest red-shift in COBeNH are consequences of the smallest OC←BeSO4 π-back-donation and the largest CO←BeNH π-back-donation, respectively.
Co-reporter:Zhuang Wu, Jian Xu, Qifan Liu, Xuelin Dong, Dingqing Li, Nicole Holzmann, Gernot Frenking, Tarek Trabelsi, Joseph S. Francisco and Xiaoqing Zeng
Physical Chemistry Chemical Physics 2017 - vol. 19(Issue 25) pp:NaN16720-16720
Publication Date(Web):2017/05/31
DOI:10.1039/C7CP02774E
A biologically relevant reactive sulfur species (RSS), the hypothiocyanite radical OSCN, is generated in the gas phase through flash vacuum pyrolysis (FVP) of trifluoromethyl sulfinyl cyanide CF3S(O)CN at ca. 1000 K. Upon UV light irradiation (365 nm), OSCN rearranges to novel isomers OSNC and SOCN, and further visible light irradiation (400 ± 20 nm) leads to reverse isomerization. The identification of OSCN, OSNC, and SOCN in cryogenic matrices (Ar and N2, 2.8 K) with IR spectroscopy is supported by quantum chemical calculations up to the CCSD(T)-F12/VTZ-F12 level. The potential energy surface for the interconversion of OSCN isomers and their bonding properties are computationally explored by using the CCSD(T)-F12/VTZ-F12 and EDA–NOCV methods, respectively.
Co-reporter:Bin Li, Subrata Kundu, Hongping Zhu, Helena Keil, Regine Herbst-Irmer, Dietmar Stalke, Gernot Frenking, Diego M. Andrada and Herbert W. Roesky
Chemical Communications 2017 - vol. 53(Issue 17) pp:NaN2546-2546
Publication Date(Web):2017/01/27
DOI:10.1039/C7CC00325K
The reaction of LAl: (L = HC[C(Me)N(Ar)]2, Ar = 2,6-iPr2C6H3) and cAAC:→AlX3 (X = Cl, I) (cAAC = cyclic alkyl amino carbene) results in strikingly asymmetric Al(II)–Al(II) compounds LAl(X)–Al(X)2–cAAC [X = Cl (2a); I (2b)]. In these dialuminum(II) compounds the two Al atoms bear different ligand environments. For a detailed insight into the structures, theoretical calculations were carried out.
Co-reporter:Kerstin Freitag, Mariusz Molon, Paul Jerabek, Katharina Dilchert, Christoph Rösler, Rüdiger W. Seidel, Christian Gemel, Gernot Frenking and Roland A. Fischer
Chemical Science (2010-Present) 2016 - vol. 7(Issue 10) pp:NaN6421-6421
Publication Date(Web):2016/06/23
DOI:10.1039/C6SC02106A
The analogy between ZnR fragments and the hydrogen radical represents a fruitful concept in organometallic synthesis. The organozinc(II) and -zinc(I) sources ZnMe2 (Me = methyl) and [Zn2Cp*2] (Cp* = pentamethylcyclopentadienyl) provide one-electron fragments ·ZnR (R = Me, Cp*), which can be trapped by transition metal complexes [LaM], yielding [Lb(ZnR)n]. The addition of the dizinc compound [Zn2Cp*2] to coordinatively unsaturated [LaM] by the homolytic cleavage of the Zn–Zn bond can be compared to the classic oxidative addition reaction of H2, forming dihydride complexes [LaM(H)2]. It has also been widely shown that dihydrogen coordinates under preservation of the H–H bond in the case of certain electronic properties of the transition metal fragment. The σ-aromatic triangular clusters [Zn3Cp*3]+ and [Zn2CuCp*3] may be regarded as the first indication of this so far unknown, side-on coordination mode of [Zn2Cp*2]. With this background in mind the question arises if a series of complexes featuring the Zn2M structural motif can be prepared exhibiting a (more or less) intact Zn–Zn interaction, i.e. di-zinc complexes which are analogous to non-classical dihydrogen complexes of the Kubas type. In order to probe this idea, a series of interrelated organozinc nickel and palladium complexes and clusters were synthesized and characterized as model compounds: [Ni(ZnCp*)(ZnMe)(PMe3)3] (1), [Ni(ZnCp*)2(ZnMe)2(PMe3)2] (2), [{Ni(CNtBu)2(μ2-ZnCp*)(μ2-ZnMe)}2] (3), [Pd(ZnCp*)4(CNtBu)2] (4) and [Pd3Zn6(PCy3)2(Cp*)4] (5). The dependence of Zn⋯Zn interactions as a function of the ligand environments and the metal centers was studied. Experimental X-ray crystallographic structural data and DFT calculations support the analogy between dihydrogen and dizinc transition metal complexes.
Co-reporter:Rajendra S. Ghadwal, Dennis Rottschäfer, Diego M. Andrada, Gernot Frenking, Christian J. Schürmann and Hans-Georg Stammler
Dalton Transactions 2017 - vol. 46(Issue 24) pp:NaN7799-7799
Publication Date(Web):2017/05/22
DOI:10.1039/C7DT01199G
The synthesis and characterization of the N-heterocyclic carbene (NHC) stabilized dichlorosilylene Group 6 metal complexes {(IPr)SiCl2}W(CO)5 (3-W), {(IPr)SiCl2}2Cr(CO)4 (4-Cr), and {(IPr)SiCl2}2W(CO)4 (4-W) (IPr = 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene) are reported. Treatment of 3-W with CsOH in the presence of IPr leads to the formation of an abnormal-NHC (aNHC) metal complex (aIPrH)W(CO)5 (6-W) (aIPrH = 1,3-bis(2,6-diisopropylphenyl)imidazol-4-ylidene), unveiling an unprecedented normal-to-abnormal transformation route of an NHC. DFT calculations support the proposed mechanism that involves CsOH mediated deprotonation of the IPr-backbone of 3-W to yield a ditopic carbanionic-NHC (dcNHC) complex 5a-W. Subsequent 1,4-migration of the W(CO)5 moiety and hydrolysis of the unmasked SiCl2 rationalize the formation of 6-W. The desired H2O molecule is generated in the initial step on deprotonation of IPr with CsOH. In contrast to the literature precedents, the calculations indicate that the abnormal complex 6-W is 13.5 kcal mol−1 thermodynamically higher in energy than the normal counterpart (IPr)W(CO)5 (8-W). Interestingly, as the aNHC-compounds reported so far are more stable than their normal counterparts, this finding showcases an opposite trend. Moreover, reaction pathways to the synthesized and related complexes have been investigated by DFT calculations.
Co-reporter:Subrata Kundu, Prinson P. Samuel, Anna Luebben, Diego M. Andrada, Gernot Frenking, Birger Dittrich and Herbert W. Roesky
Dalton Transactions 2017 - vol. 46(Issue 24) pp:NaN7952-7952
Publication Date(Web):2017/05/26
DOI:10.1039/C7DT01796K
The cyclic alkyl(amino) carbene (cAAC) [:C{N-C6H3(2,6-IPr2)}(CMe2)2CH2] stabilized MeGeGeMe has been isolated in the molecular form with composition (cAAC)MeGe–GeMe(cAAC) (1) at room temperature. Compound 1 was synthesized from the reduction of MeGeCl3 using three equivalents of KC8 in the presence of one equivalent of cAAC. The corresponding silicon compound (cAAC)MeSi–SiMe(cAAC) (2) was also prepared. Compounds 1 and 2 are the first examples of REER compounds (E = Ge, Si) carrying the smallest organic group. Furthermore the structures of compounds 1 and 2 have been investigated by using theoretical methods. The theoretical analysis of the structure of 1 is in agreement with the formation as an unprecedented carbene stabilized bis-germylene whereas compound 2 can be equally described as carbene stabilized bis-silylene with coordinate bonds as with classical double bonds of a 2,3-disila-1,3-butadiene. The compounds were also characterized by X-ray crystallography.
Co-reporter:Chelladurai Ganesamoorthy, Sinah Loerke, Christian Gemel, Paul Jerabek, Manuela Winter, Gernot Frenking and Roland A. Fischer
Chemical Communications 2013 - vol. 49(Issue 28) pp:NaN2860-2860
Publication Date(Web):2013/01/22
DOI:10.1039/C3CC38584A
Compounds Cp*AlH2 (1) and Cp*2AlH (2) reductively eliminate Cp*H in benzene or toluene under reflux conditions to give Al(s) and AlCp*, respectively.
Co-reporter:Yan Li, Hongping Zhu, Diego M. Andrada, Gernot Frenking and Herbert W. Roesky
Chemical Communications 2014 - vol. 50(Issue 35) pp:NaN4630-4630
Publication Date(Web):2014/03/10
DOI:10.1039/C4CC00912F
An interesting aminosilanetrithiol RSi(SH)3 (R = N(SiMe3)-2,6-iPr2C6H3) has been prepared by the reaction of lithium aminosilanetrithiolate {RSi[SLi(THF)]3}2 with MeCOOH. Theoretical calculations indicate that the LP(N) → σ*(Si–S) and LP(S) → σ*(Si–S) electron donations remarkably contribute to the stabilization of the Si(SH)3 part of the molecule. RSi(SH)3 is the first example of a stable molecule containing three SH groups attached to one element.
Co-reporter:Rajendra S. Ghadwal, Ramachandran Azhakar, Herbert W. Roesky, Kevin Pröpper, Birger Dittrich, Catharina Goedecke and Gernot Frenking
Chemical Communications 2012 - vol. 48(Issue 66) pp:NaN8188-8188
Publication Date(Web):2012/07/02
DOI:10.1039/C2CC32887A
Formyl chloride (H(Cl)CO) is unstable at room temperature and decomposes to HCl and CO. Silicon analogue of formyl chloride, silaformyl chloride IPr·SiH(Cl)O·B(C6F5)3 (3) (IPr = 1,3-bis(2,6-diisopropyl-phenyl)imidazol-2-ylidene), was stabilized by Lewis donor–acceptor ligands. Compound 3 is not only the first stable acyclic silacarbonyl compound but also the first silacarbonyl halide reported so far.
Co-reporter:Cameron Jones, Anastas Sidiropoulos, Nicole Holzmann, Gernot Frenking and Andreas Stasch
Chemical Communications 2012 - vol. 48(Issue 79) pp:NaN9857-9857
Publication Date(Web):2012/08/16
DOI:10.1039/C2CC35228A
Reduction of an N-heterocyclic carbene (NHC) adduct of SnCl2, viz. [(IPr)SnCl2] (IPr = :C{N(Dip)C(H)}2; Dip = 2,6-diisopropylphenyl), with a magnesium(I) dimer, has afforded the first NHC complex of a row 5 element in its diatomic form, [(IPr)SnSn(IPr)]; a computational analysis of the complex indicates that it comprises a singlet state, doubly bonded tin(0) fragment, :SnSn:, datively bonded by two NHC ligands.
Co-reporter:Ralf Tonner and Gernot Frenking
Chemical Communications 2008(Issue 13) pp:
Publication Date(Web):
DOI:10.1039/B717511F
Co-reporter:Nicole Holzmann, Markus Hermann and Gernot Frenking
Chemical Science (2010-Present) 2015 - vol. 6(Issue 7) pp:NaN4094-4094
Publication Date(Web):2015/06/01
DOI:10.1039/C5SC01504A
Quantum chemical calculations of the compound B2(NHCMe)2 and a thorough examination of the electronic structure with an energy decomposition analysis provide strong evidence for the appearance of boron–boron triple bond character. This holds for the model compound and for the isolated diboryne B2(NHCR)2 of Braunschweig which has an even slightly shorter B–B bond. The bonding situation in the molecule is best described in terms of NHCMe→B2←NHCMe donor–acceptor interactions and concomitant π-backdonation NHCMe←B2→NHCMe which weakens the B–B bond, but the essential features of a triple bond are preserved. An appropriate formula which depicts both interactions is the sketch NHCMe⇄BB⇄NHCMe. Calculations of the stretching force constants FBB which take molecules that have genuine single, double and triple bonds as references suggest that the effective bond order of B2(NHCMe)2 has the value of 2.34. The suggestion by Köppe and Schnöckel that the strength of the boron–boron bond in B2(NHCH)2 is only between a single and a double bond is repudiated. It misleadingly takes the force constant FBB of OBBO as the reference value for a B–B single bond which ignores π bonding contributions. The alleged similarity between the B–O bonds in OBBO and the B–C bonds in B2(NHCMe)2 is a mistaken application of the principle of isolable relationship.
Co-reporter:Terrance J. Hadlington, Markus Hermann, Gernot Frenking and Cameron Jones
Chemical Science (2010-Present) 2015 - vol. 6(Issue 12) pp:NaN7257-7257
Publication Date(Web):2015/09/22
DOI:10.1039/C5SC03376D
Reactions of the solution stable, two-coordinate hydrido-tetrylenes, :E(H)(L†) (E = Ge or Sn; L† = –N(Ar†)(SiPri3); Ar† = C6H2{C(H)Ph2}2Pri-2,6,4), with a variety of unactivated cyclic and acyclic alkenes, and one internal alkyne, lead to the rapid and regiospecific hydrometallation of the unsaturated substrate at ambient temperature. The products of the reactions, [L†E(C2H4R)] (E = Ge or Sn, R = H, Ph or But), [L†E{CH(CH2)3(CH2)n}] (E = Ge, n = 1, 2 or 3; E = Sn, n = 1) and [L†E{C(Ph)C(H)(Me)}], include the first structurally characterised examples of two-coordinate amido/alkyl germylenes and stannylenes. The cycloalkene hydrometallation reactions are cleanly reversible under ambient conditions, a process which computational and experimental van't Hoff analyses suggest proceeds via β-hydride elimination from the metal coordinated cycloalkyl ligand. Similarly, the reactions of :Ge(H)(L†) with 1,5-cyclooctadiene and 2-methyl-2-butene, both likely proceed via β-hydride elimination processes, leading to the clean isomerisation of the alkene involved, and its subsequent hydrogermylation, to give [L†Ge(2-cyclooctenyl)] and [L†Ge{C2H4C(H)Me2}], respectively. Reactions of [L†GeEt] and [L†Ge(C5H9)] with the protic reagents, HCl, NH3 and EtOH, lead to oxidative addition to the germanium(II) centre, and formation of the stable chiral germanium(IV) complexes, [L†Ge(C5H9)(H)Cl] and [L†Ge(Et)(H)R] (R = NH2 or OEt). In contrast, related reactions between [L†SnEt] and ButOH or TEMPOH (TEMP = 2,2,6,6-tetramethylpiperidinyl) proceed via ethane elimination, affording the tin(II) products, [L†SnR] (R = OBut or OTEMP). In addition, the oxidation of [L†Ge(C6H11)] and [L†Sn(C2H4But)] with O2 yields the oxo-bridged metal(IV) dimers, [{L†(C6H11)Ge(μ-O)}2] and [{L†(ButC2H4)Sn(μ-O)}2], respectively.
Co-reporter:Shannon A. Couchman, Nicole Holzmann, Gernot Frenking, David J. D. Wilson and Jason L. Dutton
Dalton Transactions 2013 - vol. 42(Issue 32) pp:NaN11384-11384
Publication Date(Web):2013/04/02
DOI:10.1039/C3DT50563D
A theoretical study of compounds containing Be in the +1 or 0 oxidation state has been carried out. The molecules considered containing Be in the +1 oxidation state are analogues of the important Mg(I)–Mg(I) dimer supported by the β-diketiminate ligand. The molecules in the 0 oxidation state are NHC supported compounds analogous to “molecular allotropes” which has recently become a topic of importance in p-block chemistry. In this case, our results demonstrate that the Be(0) complexes are far more stable than the analogous Mg(0) complexes, highlighting the opportunities afforded in Be chemistry, despite the challenges presented by the toxicity of Be compounds.
Co-reporter:Catharine Esterhuysen and Gernot Frenking
Dalton Transactions 2013 - vol. 42(Issue 37) pp:NaN13356-13356
Publication Date(Web):2013/05/23
DOI:10.1039/C3DT32872D
DFT calculations using BP86 in conjunction with the SVP and TZVPP basis sets as well as ab initio calculations at SCS-MP2 have been carried out for six dicoordinated carbon molecules CLL′ where L is a fluorenyl carbene while L′ is a phosphine PH3 (1) or PPh3 (2) or a carbene, i.e. NHCMe (3), benzannulated NHCMe (4), cycloheptatrienylidene (5) and benzannulated cycloheptatrienylidene (6). The complexes of these compounds with one and two AuCl moieties were also calculated. The monoaurated adducts of 1–4 have the AuCl fragment η1 coordinated to the central carbon atom. The complexes 5(AuCl) and 6(AuCl) have AuCl η2 bonded across a CC double bond. Three different bonding modes are found as energy minima for the diaurated species LL′C-(AuCl)2. The AuCl fragments are found to be either both coordinated η1, both coordinated η2 across double bonds, or a combination of the two. According to the electronic structure analysis of the free compounds, 1 and 2 might best be classified as carbenes, 3 and 4 as bent allenes while 5 and 6 are typical allenes. The complexation with AuCl reveals that 1–4 may exhibit chemical behaviour which is typical for carbones and thus, they may be termed “hidden carbones”. The AuCl complexes show that compounds 5 and 6 are classical allenes.