Jose M. Goicoechea

Find an error

Name:
Organization: University of Oxford , England
Department: Department of Chemistry
Title: (PhD)

TOPICS

Co-reporter:E. N. Faria;A. R. Jupp;J. M. Goicoechea
Chemical Communications 2017 vol. 53(Issue 52) pp:7092-7095
Publication Date(Web):2017/06/27
DOI:10.1039/C7CC01285C
We describe the reactivity of the 2-phosphaethynolate anion (PCO−) towards enantiomerically pure α-amino acids (AAs) resulting in the formation of novel salts of phosphinecarboxamides bearing chiral functionalities. These transformations occurred quantitatively with all but one of the amino acids trialled (the basic amino acid arginine was found to be unreactive). The resulting ionic species can be readily protonated to afford N-(phosphanyl)carbonyl-amino acids, a novel group of amino acids bearing primary phosphine functionalities.
Co-reporter:Jordan B. Waters;Qien Chen;Thomas A. Everitt
Dalton Transactions 2017 vol. 46(Issue 36) pp:12053-12066
Publication Date(Web):2017/09/19
DOI:10.1039/C7DT02431B
The reactivity of the heavier group 15 tribromides, SbBr3 and BiBr3, towards 1,3-bis(2,6-diisopropylphenyl)-imidazol-2-ylidene (IPr) is described. These reactions quantitatively afford Lewis acid–base adducts, (IPr)EBr3 (E = Sb 1; Bi 2), which readily react with AlBr3 yielding cationic species [(IPr)EBr2]+ (E = Sb 3; Bi 4). Under thermal treatment, the N-heterocyclic carbene ligands in 1 and 2 will readily isomerise to afford the abnormally-bonded (or mesoionic) complexes (aIPr)EBr3 (E = Sb 5; Bi 6). As with 1 and 2, bromide abstraction from such compounds readily affords the cationic complexes [(aIPr)EBr2]+ (E = Sb 7; Bi 8). Finally, in an effort to elucidate the isomerisation process which allows for the conversion of 1 and 2 to the abnormally bonded systems (compounds 5 and 6), compound 1 was reacted with a further equivalent of IPr to afford the cationic species [(aIPr)2SbBr2]+ (9). This strongly suggests that the normal to abnormal isomerisation of the N-heterocylic carbene ligands in compounds 1 and 2 is mediated by the presence of free IPr. Compound [9]Br can be used to access the dicationic species [(aIPr)2SbBr]2+ (10), which we have identified spectroscopically. Single crystal X-ray structures and spectroscopic data for all compounds are discussed.
Co-reporter:Adinarayana Doddi;Michael Weinhart;Alexander Hinz;Dirk Bockfeld;Manfred Scheer;Matthias Tamm
Chemical Communications 2017 vol. 53(Issue 45) pp:6069-6072
Publication Date(Web):2017/06/01
DOI:10.1039/C7CC02628E
N-Heterocyclic carbene adducts of the parent arsinidene (AsH) were prepared by two different synthetic routes, either by reaction of As(SiMe3)3 with 2,2-difluoroimidazolines followed by desilylation or by reaction of [Na(dioxane)3.31][AsCO] with imidazolium chlorides.
Co-reporter:Gabriela Espinoza Quintero, Isabelle Paterson-Taylor, Nicholas H. Rees and Jose M. Goicoechea  
Dalton Transactions 2016 vol. 45(Issue 5) pp:1930-1936
Publication Date(Web):17 Jul 2015
DOI:10.1039/C5DT02380G
Reactions of the protonated heptaphosphide dianion, [HP7]2−, with one equivalent of E[N(SiMe3)2]2 (E = Ge, Sn, Pb) give rise to novel derivatized cluster anions [P7EN(SiMe3)2]2− (E = Ge (1), Sn (2) and Pb (3)). All three species were characterized by multi-element solution-phase NMR spectroscopy and electrospray ionization mass spectrometry. In addition, 1 and 2 were structurally authenticated by means of single crystal X-ray diffraction in [K(18-crown-6)]2[P7EN(SiMe3)2]·2py. Interestingly, while 2 appears to be indefinitely stable in solution for prolonged periods of time, the germanium-containing analogue, 1, readily decomposes at room temperature giving rise to the dimeric species [(P7Ge)2N(SiMe3)2]3− (4) and [K(18-crown-6)][N(SiMe3)2]. A low quality single crystal X-ray structure of the former allowed for the confirmation of its composition and connectivity which is consistent with the 31P NMR spectrum obtained for the anion.
Co-reporter:Andrew R. Jupp;Michael B. Geeson;John E. McGrady
European Journal of Inorganic Chemistry 2016 Volume 2016( Issue 5) pp:639-648
Publication Date(Web):
DOI:10.1002/ejic.201501075

Abstract

A synthesis of the 2-phosphathioethynolate anion, PCS, under ambient conditions is reported. The coordination chemistry of PCO, PCS and their nitrogen-containing congeners is also explored. Photolysis of a solution of W(CO)6 in the presence of PCO [or a simple ligand displacement reaction using W(CO)5(MeCN)] affords [W(CO)5(PCO)] (1). The cyanate and thiocyanate analogues, [W(CO)5(NCO)] (2) and [W(CO)5(NCS)] (3), are also synthesised using a similar methodology, allowing for an in-depth study of the bonding properties of this family of related ligands. Our studies reveal that, in the coordination sphere of tungsten(0), the PCO anion preferentially binds through the phosphorus atom in a strongly bent fashion, while NCO and NCS coordinate linearly through the nitrogen atom. Reactions between PCS and W(CO)5(MeCN) similarly afford [W(CO)5(PCS)]; however, due to the ambidentate nature of the anion, a mixture of both the phosphorus- and sulfur-bonded complexes (4a and 4b, respectively) is obtained. It was possible to establish that, as with PCO, the PCS ion also coordinates to the metal centre in a bent fashion.

Co-reporter:Dr. Alexer Hinz ;Dr. Jose M. Goicoechea
Angewandte Chemie International Edition 2016 Volume 55( Issue 30) pp:8536-8541
Publication Date(Web):
DOI:10.1002/anie.201602310

Abstract

The synthesis and isolation of the 2-arsaethynolate anion, AsCO, and its subsequent reactivity towards heteroallenes is reported. Reactions with ketenes and carbodiimides afford four-membered anionic heterocycles in formal [2+2] cycloaddition reactions. By contrast, reaction with an isocyanate yielded a 1,4,2-diazaarsolidine-3,5-dionide anion and the unprecedented cluster anions As102− and As124−. These preliminary reactivity studies hint at the enormous potential synthetic utility of this novel anion, which may be employed as an arsenide (As) source.

Co-reporter:Dr. Alexer Hinz ;Dr. Jose M. Goicoechea
Angewandte Chemie International Edition 2016 Volume 55( Issue 30) pp:
Publication Date(Web):
DOI:10.1002/anie.201683061
Co-reporter:Dr. Alexer Hinz ;Dr. Jose M. Goicoechea
Angewandte Chemie 2016 Volume 128( Issue 30) pp:8678-8683
Publication Date(Web):
DOI:10.1002/ange.201602310

Abstract

The synthesis and isolation of the 2-arsaethynolate anion, AsCO, and its subsequent reactivity towards heteroallenes is reported. Reactions with ketenes and carbodiimides afford four-membered anionic heterocycles in formal [2+2] cycloaddition reactions. By contrast, reaction with an isocyanate yielded a 1,4,2-diazaarsolidine-3,5-dionide anion and the unprecedented cluster anions As102− and As124−. These preliminary reactivity studies hint at the enormous potential synthetic utility of this novel anion, which may be employed as an arsenide (As) source.

Co-reporter:Dr. Alexer Hinz ;Dr. Jose M. Goicoechea
Angewandte Chemie 2016 Volume 128( Issue 30) pp:
Publication Date(Web):
DOI:10.1002/ange.201683061
Co-reporter:Jordan B. Waters, Jose M. Goicoechea
Coordination Chemistry Reviews 2015 Volumes 293–294() pp:80-94
Publication Date(Web):15 June 2015
DOI:10.1016/j.ccr.2014.09.020
•A comprehensive review of the chemistry of ditopic carbanionic N-heterocyclic carbenes (dcNHCs) is provided.•The isolation of dcNHCs as alkali-metal salts is described.•The reactivity of alkali metal salts of dNHCs is reviewed.The chemistry of N-heterocyclic carbenes, particularly imidazol-2-ylidenes, has been exhaustively documented in the literature. Amongst the many bond activation reactions available for such species (CH, CC and NC), it has recently become apparent that deprotonation of the alkenic backbone can give rise to ditopic carbanionic carbenes, i.e. anionic species capable of bonding to Lewis acidic centres via two different carbon sites (C2 and C4 or C4 and C5). In this article, we aim to provide a comprehensive review of the chemistry of such species, with a particular focus on their isolation as alkali-metal salts and their subsequent reactivity.
Co-reporter:Andrew R. Jupp;Gemma Trott;Éléonore PayendelaGarerie;James D. G. Holl;Dr. Duncan Carmichael;Dr. Jose M. Goicoechea
Chemistry - A European Journal 2015 Volume 21( Issue 22) pp:8015-8018
Publication Date(Web):
DOI:10.1002/chem.201501174

Abstract

We demonstrate that the synthesis of new N-functionalized phosphinecarboxamides is possible by reaction of primary and secondary amines with PCO in the presence of a proton source. These reactions proceed with varying degrees of success, and although primary amines generally afford the corresponding phosphinecarboxamides in good yields, secondary amines react more sluggishly and often give rise to significant decomposition of the 2-phosphaethynolate precursor. Of the new N-derivatized phosphinecarboxamides available, PH2C(O)NHCy (Cy=cyclohexyl) can be obtained in sufficiently high yields to allow for the exploration of its Brønsted acidity. Thus, deprotonating PH2C(O)NHCy with one equivalent of potassium bis(trimethylsilyl)amide (KHMDS) gave the new phosphide [PHC(O)NHCy]. In contrast, deprotonation with half of an equivalent gives rise to [P{C(O)NHCy}2] and PH3. These phosphides can be employed to give new phosphines by reactions with electrophiles, thus demonstrating their enormous potential as chemical building blocks.

Co-reporter:Dr. Thomas P. Robinson ;Dr. Jose M. Goicoechea
Chemistry - A European Journal 2015 Volume 21( Issue 15) pp:5727-5731
Publication Date(Web):
DOI:10.1002/chem.201500628

Abstract

We report that the 2-phosphaethynolate anion (PCO) reacts with propargylamines in the presence of a proton source to afford novel N-derivatized phosphinecarboxamides bearing alkyne functionalities. Deprotonation of these species gives rise to novel five- and six-membered anionic heterocycles resulting from intramolecular nucleophilic attack of the resulting phosphide at the alkyne functionality (via 5-exo-dig or 6-endo-dig cyclizations, respectively). The nature of the substituents on the phosphinecarboxamide can be used to influence the outcome of these reactions. This strategy represents a unique approach to phosphorus-containing heterocylic systems that are closely related to known organic molecules with interesting bio-active properties.

Co-reporter:Dr. Thomas P. Robinson;Dr. Michael J. Cowley;Dr. David Scheschkewitz;Dr. Jose M. Goicoechea
Angewandte Chemie International Edition 2015 Volume 54( Issue 2) pp:683-686
Publication Date(Web):
DOI:10.1002/anie.201409908

Abstract

The reactivity of the 2-phosphaethynolate anion (PCO) towards a cyclic trisilene (cSi3(Tip)4) is reported. The result is the net activation of the PC and SiSi multiple bonds of the precursors affording a heteroatomic bicyclo[1.1.1]pentan-2-one analogue ([P(CO)Si3(Tip)4]; 1). This reaction can be interpreted as the formal addition of a phosphide and a carbonyl across the SiSi double bond. Photolytic decarbonylation of 1 results in the incorporation of the phosphide vertex into the cyclotrisilene scaffold, yielding a congener of the cyclobutene anion with considerable allylic character.

Co-reporter:Robert S. P. Turbervill and Jose M. Goicoechea
Chemical Reviews 2014 Volume 114(Issue 21) pp:10807
Publication Date(Web):October 21, 2014
DOI:10.1021/cr500387w
Co-reporter:M. B. Geeson, A. R. Jupp, J. E. McGrady and J. M. Goicoechea  
Chemical Communications 2014 vol. 50(Issue 82) pp:12281-12284
Publication Date(Web):27 Aug 2014
DOI:10.1039/C4CC06094F
We describe the coordination chemistry of the primary phosphine PH2C(O)NH2 (phosphinecarboxamide) towards group 6 transition-metals. Experimental and theoretical studies reveal that this novel species has comparable electronic properties to PH3.
Co-reporter:Rebecca A. Musgrave, Robert S. P. Turbervill, Mark Irwin, Radovan Herchel and Jose M. Goicoechea  
Dalton Transactions 2014 vol. 43(Issue 11) pp:4335-4344
Publication Date(Web):16 Oct 2013
DOI:10.1039/C3DT52638K
Reaction of dimesityliron(II) (Fe2(mes)4) with the N-heterocyclic carbenes 1,3-bis(2,6-diisopropylphenyl)-imidazol-2-ylidene (IPr) and 1,3-bis(2,6-dimethylphenyl)hexahydropyrimidin-2-ylidene (6-Xyl) afforded the novel trigonal planar complexes [Fe(IPr)(mes)2] (1) and [Fe(6-Xyl)(mes)2] (2), respectively. Both species were structurally characterized by single crystal X-ray diffraction and display structures and magnetic responses consistent with a quintet ground state (S = 2). Reaction of 1 with KC8 in THF afforded K+ salts of the anionic complex [{:C[N(2,6-iPr2C6H3)]2(CH)C}2Fe(mes)]− (3) and the homoleptic organometallic anion [Fe(mes)3]− (4). By contrast, reduction of 2 resulted in extensive decomposition and intractable product mixtures. Complex 3 is coordinated by two ditopic carbanionic carbenes via the C4/C5 position while the C2 position retains unquenched carbenic character and remains vacant for further coordination. This was corroborated by reacting solutions of 3 with one and two equivalents of triethylaluminium (AlEt3) which resulted in the formation of [{Et3Al:C[N(2,6-iPr2C6H3)]2(CH)C}{:C[N(2,6-iPr2C6H3)]2(CH)C}Fe(mes)]− (5) and [{Et3Al:C[N(2,6-iPr2C6H3)]2(CH)C}2Fe(mes)]− (6), respectively. Both of these species were structurally characterized as [K(2,2,2-crypt)]+ salts.
Co-reporter:Jordan B. Waters and Jose M. Goicoechea  
Dalton Transactions 2014 vol. 43(Issue 38) pp:14239-14248
Publication Date(Web):06 May 2014
DOI:10.1039/C4DT00954A
Reaction of the lithiated N-heterocyclic carbene [:C[N(2,6-iPr2C6H3)]2(CH)CLi]n (LiIPr) with KOtBu in diethylether (Et2O) afforded the novel organo-potassium compound [:C[N(2,6-iPr2C6H3)]2(CH)CK(THF)2] (KIPr·2THF). Both LiIPr and KIPr can be interpreted as ditopic carbanionic carbenes (or alkali metal salts of anionic “dicarbenes”) and are interesting precursors for the synthesis of novel metal complexes bearing carbanionic carbenes as ligands. Reaction of KIPr with M[N(SiMe3)2]2 (M = Zn, Sn) afforded salts of the anionic three coordinate complexes [M{C(CH)[N(2,6-iPr2C6H3)]2C:}{N(SiMe3)2}2]− (M = Zn (1) and Sn (2)). Contrasting reactivity was observed for the other group 14 bis-amide compounds M[N(SiMe3)2]2 (M = Ge, Pb), which initially appear to yield analogous 1:1 complexes (M = Ge (3) and Pb (4)), however over time give rise to compounds bearing two ditopic carbanionic carbenes ([M{C(CH)[N(2,6-iPr2C6H3)]2C:}2{N(SiMe3)2}]−; Ge (5) and Pb (6)) and the tris-amide anions ([M{N(SiMe3)3}]−), presumably via a Schlenk-type equilibrium. Compounds 5 and 6 can be directly synthesized by reacting M[N(SiMe3)2]2 (M = Ge, Pb) with two equivalents of KIPr, respectively.
Co-reporter:Robert S. P. Turbervill
European Journal of Inorganic Chemistry 2014 Volume 2014( Issue 10) pp:1660-1668
Publication Date(Web):
DOI:10.1002/ejic.201301011

Abstract

By following a similar procedure to that employed for the synthesis of [HP7]2– (1), the synthesis and characterization of the hydrogenheptaarsenide dianion [HAs7]2– (2) was achieved. Building on previous research that has shown that [HP7]2– can react with carbodiimides to yield amidine-functionalized cluster anions [P7C(NHR)(NR)]2–, we have found that the analogous hydroarsination reactions are also possible by reaction of 2 with RC=N=CR to yield [As7C(NHR)(NR)]2– [R = Dipp (3), iPr (4) and Cy (5); Dipp = 2,6-diisopropylphenyl]. Furthermore, we also demonstrate that such hydropnictination reactions can be extended to other heteroallenes such as isocyanates (O=C=NAd; Ad = adamantyl) to afford the amide-derivatized cluster anions [E7C(CO)(NHAd)]2– [E = P (6) and As (7)]. All new compounds were characterized by multinuclear NMR spectroscopy, elemental analyses and electrospray mass spectrometry. Clusters 2, 4, 6 and 7 were also characterized by single-crystal X-ray diffraction of [K(18-crown-6)]2[2], [K(18-crown-6)]2[4], [K(18-crown-6)]2[6]·py (py = pyridine) and [K(18-crown-6)]2[7]·py. Crystal structures confirming composition and connectivity were also obtained for anions 3 and 5.

Co-reporter:Andrew R. Jupp
Journal of the American Chemical Society 2013 Volume 135(Issue 51) pp:19131-19134
Publication Date(Web):December 6, 2013
DOI:10.1021/ja4115693
Reactions of the 2-phosphaethynolate anion (PCO–, 1) with ammonium salts quantitatively yielded phosphinecarboxamide (PH2C(O)NH2, 2). The molecular structure and chemical properties of 2 were studied by single-crystal X-ray diffraction and multielement NMR spectroscopy. This phosphorus-containing analogue of urea is a rare example of an air-stable primary phosphine.
Co-reporter:Gabriela Espinoza-Quintero ; Jack C. A. Duckworth ; William K. Myers ; John E. McGrady
Journal of the American Chemical Society 2013 Volume 136(Issue 4) pp:1210-1213
Publication Date(Web):December 30, 2013
DOI:10.1021/ja411280v
The 12-vertex endohedral cluster [Ru@Ge12]3– reveals an unprecedented D2d-symmetric 3-connected polyhedral geometry. The structure contrasts dramatically with the known deltahedral or approximately deltahedral geometries of [M@Pb12]2– (M = Ni, Pd, Pt) and [Mn@Pb12]3– and is a result of extensive delocalization of electron density from the transition-metal center onto the cage.
Co-reporter:Robert S. P. Turbervill and Jose M. Goicoechea
Inorganic Chemistry 2013 Volume 52(Issue 9) pp:5527-5534
Publication Date(Web):April 22, 2013
DOI:10.1021/ic400448x
Reactions between anionic heptaphosphide clusters ([P7]3–/[HP7]2–) and 2-ethynylpyridine yielded the 4-(2′-pyridyl)-1,2,3-triphospholide anion ([P3C2H(2-C5H4N)]−; 1). This species was isolated as a compositionally pure [K(2,2,2-crypt)]+ salt in moderate yields. Preliminary coordination studies of 1 toward Mo(CO)6 or Mo(py)3(CO)3 (py = pyridine) afforded the diamagnetic piano-stool complex [{η5-P3C2H(2-C5H4N)}Mo(CO)3]− (2). By contrast, reaction of 1 with Mo(COD)(CO)4 (COD = 1,5-cyclooctadiene) yielded [{κ2P,N-P3C2H(2-C5H4N)}Mo(CO)4]− (3) which readily loses a carbonyl on heating to give 2. Reaction of 2 with Mo(COD)(CO)4 afforded the bimetallic system [{μ:η5,κ2P,N-P3C2H(2-C5H4N)}{Mo(CO)3}{Mo(CO)4}]− (4).
Co-reporter:Tobias Krämer, Jack C. A. Duckworth, Matthew D. Ingram, Binbin Zhou, John E. McGrady and Jose M. Goicoechea  
Dalton Transactions 2013 vol. 42(Issue 34) pp:12120-12129
Publication Date(Web):08 Apr 2013
DOI:10.1039/C3DT50643F
The synthesis of a new endohedral ten-vertex Zintl ion cluster, [Fe@Sn10]3−, isoelectronic with [Fe@Ge10]3−, is reported. In an attempt to place this new cluster within the context of the known structural chemistry of the M@E10 family (M = transition metal, E = main group element), we have carried out a detailed electronic structure analysis of the different structural types: viz bicapped square antiprismatic ([Ni@Pb10]2−, [Zn@In10]8−), tetra-capped trigonal prismatic ([Ni@In10]10−) and the remarkable pentagonal prismatic [Fe@Ge10]3− and [Co@Ge10]3−. We establish that the structural trends can be interpreted in terms of a continuum of effective electron counts at the E10 cage, ranging from electron deficient (<4n + 2) in [Ni@In10]10− to highly electron rich (>4n + 2) in [Fe@Ge10]3−. The effective electron count differs from the total valence electron count in that it factors in the increasingly active role of the metal d electrons towards the left of the transition series. The preference for a pentagonal prismatic geometry in [Fe@Ge10]3− emerges as a natural consequence of backbonding to the cage from four orthogonal 3d orbitals of the low-valent metal ion. Our calculations suggest that the new [Fe@Sn10]3− cluster should also exhibit structural consequences of backbonding from the metal to the cage, albeit to a less extreme degree than in its Ge analogue. The global minimum lies on a very flat surface connecting D4d, C2v and C3v-symmetric minima, suggesting a very plastic structure that may be easily deformed by the surrounding crystal environment. If so, then this provides a new and quite distinct structural type for the M@E10 family.
Co-reporter:Jordan B. Waters, Robert S. P. Turbervill, and Jose M. Goicoechea
Organometallics 2013 Volume 32(Issue 18) pp:5190-5200
Publication Date(Web):September 9, 2013
DOI:10.1021/om400728u
Reaction of the N-heterocyclic carbene 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene (IPr) with M(mes)2 (M = Zn, Cd) in diethyl ether afforded the Lewis acid–base adducts [M(IPr)(mes)2] (M = Zn (1), Cd (2)) in quantitative yields. An analogous reaction between Hg(mes)2 and IPr failed to form the desired 1:1 adduct, [Hg(IPr)(mes)2], as evidenced by NMR spectroscopy. Reduction of 1 and 2 with KC8 afforded K+ salts of the anionic complexes [{:C[N(2,6-iPr2C6H3)]2(CH)C}2M(mes)]− (M = Zn (3), Cd (4)). By contrast, reduction of a THF solution of a mixture of Hg(mes)2 and IPr gave rise to the homoleptic anionic species [{:C[N(2,6-iPr2C6H3)]2(CH)C}3Hg]− (5). Species 3–5 display abnormally bonded anionic N-heterocyclic “dicarbene” ligands (or ditopic carbanionic carbenes) in which IPr has been deprotonated at the C4/C5 position. The vacant C2 atoms retain carbenic character, allowing for further coordination to Lewis acids. This was demonstrated by reaction of 3 with H3B:SMe2, AlEt3, and CO2 (in the presence of the appropriate cation-sequestering agents), which afforded salts of the [{LA:C[N(2,6-iPr2C6H3)]2(CH)C}2Zn(mes)]− anions (LA = BH3 (6), AlEt3 (7), and CO2 (8)).
Co-reporter:Robert S. P. Turbervill, Andrew R. Jupp, Phillip S. B. McCullough, Doruk Ergöçmen, and Jose M. Goicoechea
Organometallics 2013 Volume 32(Issue 7) pp:2234-2244
Publication Date(Web):March 26, 2013
DOI:10.1021/om4001296
We have investigated the chemical reactivity of heptaatomic anionic clusters of the group 15 elements ([E7]3–/[HE7]2–, E = P, As) toward the symmetric and asymmetrically substituted alkynes diphenylacetylene and phenylacetylene. The results reported herein, alongside a previous report on the reactivity of such clusters toward acetylene, describe a versatile route by which to access otherwise elusive 1,2,3-tripnictolide anions of the general formula [E3C2RR′]− (R = R′ = H, E = P (1), As (2); R = R′ = C6H5, E = P (3) As (4); R = H, R′ = C6H5, E = P (5), As (6)). These species can be isolated as [K(18-crown-6)]+ or [K(2,2,2-crypt)]+ salts. All anions were characterized by multielement NMR spectroscopy and electrospray mass spectrometry. In addition, single-crystal X-ray diffraction structures of the novel species [K(18-crown-6)(THF)2][3], [K(2,2,2-crypt)][4]·xTHF (x = 0, 0.5), and [K(18-crown-6)THF][6] were also obtained. The chemical reactivity of these group 15 cyclopentadienyl analogues has been explored in a series of ligand displacement reactions with Mo(CO)3(L)3 (L = CO, CH3CN) to yield the complex anions [(η5-E3C2H2)Mo(CO)3]− (E = P (7), As (8)), [{η5-E3C2(C6H5)2}Mo(CO)3]− (E = P (9), As (10)), and [{η5-E3C2H(C6H5)}Mo(CO)3]− (E = P (11), As (12)).
Co-reporter:Andrew R. Jupp ;Dr. Jose M. Goicoechea
Angewandte Chemie 2013 Volume 125( Issue 38) pp:10248-10251
Publication Date(Web):
DOI:10.1002/ange.201305235
Co-reporter:Andrew R. Jupp ;Dr. Jose M. Goicoechea
Angewandte Chemie International Edition 2013 Volume 52( Issue 38) pp:10064-10067
Publication Date(Web):
DOI:10.1002/anie.201305235
Co-reporter:Caroline M. Knapp, B. H. Westcott, M. A. C. Raybould, J. E. McGrady and J. M. Goicoechea  
Chemical Communications 2012 vol. 48(Issue 100) pp:12183-12185
Publication Date(Web):01 Nov 2012
DOI:10.1039/C2CC36888A
Reaction of an ethylenediamine solution of K3As7 with the low-valent, low-coordinate cobalt(II) complex [Co(mes)2(PEt2Ph)2] yielded the novel dianionic species [Co(η3-As3){η4-As4(mes)2}]2− (1). The [η4-As4(mes)2]2− moiety present in 1 is a rare example of a group 15 analogue of a butadienediide.
Co-reporter:Robert S. P. Turbervill and Jose M. Goicoechea  
Chemical Communications 2012 vol. 48(Issue 10) pp:1470-1472
Publication Date(Web):16 Jun 2011
DOI:10.1039/C1CC12089A
Reaction of the hydrogenheptaphosphide dianion, [HP7]2−, with one equivalent of bis(2,6-diisopropylphenyl)carbodiimide yielded the exo-functionalized cluster [P7C(NDipp)(NHDipp)]2− (Dipp = 2,6-diisopropylphenyl). This species was characterized by single crystal X-ray diffraction, multielement NMR spectroscopy and electrospray mass-spectrometry. DFT calculations on the dianion were also conducted.
Co-reporter:Robert S. P. Turbervill and Jose M. Goicoechea  
Chemical Communications 2012 vol. 48(Issue 49) pp:6100-6102
Publication Date(Web):23 Apr 2012
DOI:10.1039/C2CC32244G
Dimethylformamide solutions of K3E7 (E = P, As) react with acetylene yielding the 1,2,3-tripnictolide anions [E3C2H2]− (R = P (1), As (2)). Preliminary studies have shown that 1 and 2 displace labile ligands in [Ru(COD){η3-CH3C(CH2)2}2] (COD = 1,5-cyclooctadiene) to yield the novel complexes [Ru(η5-E3C2H2){CH3C(CH2)2}2}]− (E = P (3), As (4)).
Co-reporter:Mark Irwin, Laurence R. Doyle, Tobias Krämer, Radovan Herchel, John E. McGrady, and Jose M. Goicoechea
Inorganic Chemistry 2012 Volume 51(Issue 22) pp:12301-12312
Publication Date(Web):October 30, 2012
DOI:10.1021/ic301587f
The organometallic first-row transition-metal complexes [M(2,2′-bipy)(mes)2] (M = Cr (1), Mn (2), Co (4), Ni (5); 2,2′-bipy = 2,2′-bipyridine; mes = 2,4,6-Me3C6H2) were reacted with potassium and a suitable alkali-metal sequestering agent to yield salts of the anionic species [M(2,2′-bipy)(mes)2]−. The neutral parent compounds and their corresponding anionic congeners were characterized by single-crystal X-ray diffraction in [Cr(2,2′-bipy)(mes)2]·1.5C6H6, [Mn(2,2′-bipy)(mes)2], [Co(2,2′-bipy)(mes)2]·THF, [Ni(2,2′-bipy)(mes)2], [K(dibenzo-18-crown-6)·THF][Cr(2,2′-bipy)(mes)2]·2THF, [K(18-crown-6)][Mn(2,2′-bipy)(mes)2]·2THF, [K(18-crown-6)][Mn(2,2′-bipy)(mes)2]·0.67py·0.67tol, [K(2,2,2-crypt)][Co(2,2′-bipy)(mes)2], and [K(2,2,2-crypt)][Ni(2,2′-bipy)(mes)2]. These species, along with the previously reported neutral and anionic iron complexes [Fe(2,2′-bipy)(mes)2]0/– (3/3–), form a homologous series of compounds which allow for an in-depth study of the interactions between metals and ligands. Single-crystal X-ray diffraction data, DFT calculations, and various spectroscopic and magnetic measurements indicate that the anionic complexes (1––5–) can be best formulated as M(II) complexes of the 2,2′-bipyridyl radical anion. These findings complement recent studies which indicate that bond metric data from single-crystal X-ray diffraction may be employed as an important diagnostic tool in determining the oxidation states of bipyridyl ligands in transition-metal complexes.
Co-reporter:David F. Hansen, Binbin Zhou, Jose M. Goicoechea
Journal of Organometallic Chemistry 2012 s 721–722() pp: 53-61
Publication Date(Web):
DOI:10.1016/j.jorganchem.2012.05.037
Co-reporter:Caroline M. Knapp;Bethan H. Westcott;Melissa A. C. Raybould;Dr. John E. McGrady ;Dr. Jose M. Goicoechea
Angewandte Chemie International Edition 2012 Volume 51( Issue 36) pp:9097-9100
Publication Date(Web):
DOI:10.1002/anie.201203980
Co-reporter:Caroline M. Knapp;Bethan H. Westcott;Melissa A. C. Raybould;Dr. John E. McGrady ;Dr. Jose M. Goicoechea
Angewandte Chemie 2012 Volume 124( Issue 36) pp:9231-9234
Publication Date(Web):
DOI:10.1002/ange.201203980
Co-reporter:Rebecca A. Musgrave;Robert S. P. Turbervill;Mark Irwin ;Dr. Jose M. Goicoechea
Angewandte Chemie 2012 Volume 124( Issue 43) pp:10990-10993
Publication Date(Web):
DOI:10.1002/ange.201206100
Co-reporter:Rebecca A. Musgrave;Robert S. P. Turbervill;Mark Irwin ;Dr. Jose M. Goicoechea
Angewandte Chemie International Edition 2012 Volume 51( Issue 43) pp:10832-10835
Publication Date(Web):
DOI:10.1002/anie.201206100
Co-reporter:Robert S. P. Turbervill and Jose M. Goicoechea
Organometallics 2012 Volume 31(Issue 6) pp:2452-2462
Publication Date(Web):March 8, 2012
DOI:10.1021/om300072z
Direct reaction of the hydrogen heptaphosphide dianion, [HP7]2– (1), with carbodiimides RN═C═NR (R = Dipp (2,6-diisopropylphenyl), iPr, Cy) afforded the amidine-functionalized Zintl ions [P7C(NHR)(NR)]2– (R = Dipp (2), iPr (3), and Cy (4)). The bimodal activity of 1, which contains both negatively charged phosphide vertices alongside “electron-precise” phosphine-like nuclei allows for the direct hydrophosphination of the carbodiimides without need for an external proton source. Further reaction of the bis(2,6-diisopropylphenyl)amidine-functionalized cluster 2 with a proton source such as [NH4][BPh4] or [H(OEt2)2][BArF4] affords the protonated monoanionic species [HP7C(NHDipp)(NDipp)]− (5). As with 1, species 5 also has bimodal character, and additional hydrophosphination reactions are possible by reacting 5 with RN═C═NR to yield the bis-functionalized monoanions {[P7[C(NHDipp)(NDipp)][C(NHR)(NR)]}− (R = Dipp (6), iPr (7), and Cy (8)). All species were characterized by multielement NMR spectroscopy and electrospray mass spectrometry. In addition clusters 3, 5, and 6–8 were characterized by single-crystal X-ray diffraction in [K(18-crown-6)]2[3], [K(2,2,2-crypt)][5], [K(2,2,2-crypt)][6]·THF, [K(2,2,2-crypt)][7]·hex, and [K(2,2,2-crypt)][8]·0.65THF·0.35hex, respectively. Density functional theory level calculations were conducted on all anionic species to probe their electronic structure.
Co-reporter:Caroline M. Knapp, Joseph S. Large, Nicholas H. Rees and Jose M. Goicoechea  
Chemical Communications 2011 vol. 47(Issue 14) pp:4111-4113
Publication Date(Web):07 Mar 2011
DOI:10.1039/C1CC10236B
Reaction of K3P7, FeCl2, [NH4][B(C6H5)4] and 2,2,2-crypt yielded the unprecedented cluster dianion [Fe(HP7)2]2−. This species was characterized by single crystal X-ray diffraction, multielement NMR spectroscopy and electrospray mass-spectrometry. DFT calculations on the dianion were also conducted.
Co-reporter:Caroline M. Knapp, Charlotte S. Jackson, Joseph S. Large, Amber L. Thompson, and Jose M. Goicoechea
Inorganic Chemistry 2011 Volume 50(Issue 9) pp:4021-4028
Publication Date(Web):March 22, 2011
DOI:10.1021/ic102516g
Ethylenediamine (en) solutions of K3P7 and 2,2,2-crypt (4,7,13,16,21,24-hexaoxa-1,10-diazabicyclo[8.8.8]hexacosane) were reacted with the homoleptic group 11 complexes [M(nbe)3][SbF6] (M = Ag, Au; nbe = norbornene) yielding two novel cluster anions, [M2(HP7)2]2−, both of which were isolated in low crystalline yields as [K(2,2,2-crypt)]2[M2(HP7)2] (M = Ag (1) and Au (2)). Optimization of the reaction conditions by incorporation of a proton source (ammonium tetraphenylborate) and the replacement of the light-sensitive nbe adducts of silver and gold with the chloride salts MCl (M = Ag, Au) was found to greatly increase the yield and purity in which 1 and 2 were isolated. Compounds 1 and 2 were characterized by single crystal X-ray diffraction, electrospray ionization mass-spectrometry (ESI− MS), elemental analysis, and 1H and 31P NMR spectroscopy. Density functional theory (DFT) calculations on the cluster anions were also conducted.
Co-reporter:Mark Irwin ; Tobias Krämer ; John E. McGrady
Inorganic Chemistry 2011 Volume 50(Issue 11) pp:5006-5014
Publication Date(Web):May 3, 2011
DOI:10.1021/ic200241d
Addition of 1 equiv of potassium metal to a tetrahydrofuran (THF) solution of Zn2(4,4′-bipyridine)(mes)4 (1; mes =2,4,6-Me3C6H2) in the presence of 18-crown-6 (1,4,7,10,13,16-hexaoxacyclooctadecane) yielded the radical anionic species [Zn2(4,4′-bipyridine)(mes)4]•–, which was characterized by single crystal X-ray diffraction in [K(18-crown-6)(THF)2][Zn2(4,4′-bipyridine)(mes)4] (2). A similar reaction employing 2 equiv of alkali metal afforded the related complex [K(18-crown-6)]2[Zn2(4,4′-bipyridine)(mes)4] (3). The [Zn2(4,4′-bipyridine)(mes)4]n− (n = 0–2) moieties present in 1–3 are largely isostructural, yet exhibit significant structural variations which arise because of differences in their electronic structure. These species represent a homologous series of complexes in which the ligand exists in three distinct oxidation states. Structural data, spectroscopic measurements, and density functional theory (DFT) calculations are consistent with the assignment of 1, 2, and 3 as complexes of the neutral, radical anionic, and dianionic 4,4′-bipyridyl ligand, respectively. To the best of our knowledge, species 2 and 3 are the first crystallographically characterized transition metal complexes of the 4,4′-bipyridyl radical and dianion.
Co-reporter:Binbin Zhou, Tobias Krämer, Amber L. Thompson, John E. McGrady, and Jose M. Goicoechea
Inorganic Chemistry 2011 Volume 50(Issue 17) pp:8028-8037
Publication Date(Web):August 4, 2011
DOI:10.1021/ic200329m
Reaction of an ethylenediamine (en) solution of K4Pb9 and 2,2,2-crypt (4,7,13,16,21,24-hexaoxa-1,10-diazabicyclo[8.8.8]hexacosane) with a tetrahydrofuran (THF) solution of Mn3(Mes)6 (Mes = 2,4,6-trimethylphenyl) yielded the anionic cluster [Mn@Pb12]3–. This species was observed in the positive and negative ion-mode electrospray mass-spectra of the crude reaction mixture. The crystalline samples obtained from such solutions allowed us to confirm the composition of the sample as [K(2,2,2-crypt)]3[Mn@Pb12]·1.5en (1). Because of numerous issues related to crystal sample quality and crystallographic disorder a high-quality crystal structure solution could not be obtained. Despite this, however, the data collected permit us to draw reasonable conclusions about the charge and connectivity of the [Mn@Pb12]3– cluster anion. Crystals of 1 were further characterized by elemental analysis and electron paramagnetic resonance (EPR). Density Functional Theory (DFT) calculations on such a system reveal a highly distorted endohedral cluster anion, consistent with the structural distortions observed by single crystal X-ray diffraction. The cluster anions are considerably expanded compared to the 36-electron closed-shell analogue [Ni@Pb12]2– and, moreover, exhibit significant low-symmetry distortions from the idealized icosahedral (Ih) geometry that is characteristic of related endohedral clusters. Our computations indicate that there is substantial transfer of electron density from the formally Mn(−I) center to the low-lying vacant orbitals of the [Pb12]2– cage.
Co-reporter:Mark Irwin ; Rhiannon K. Jenkins ; Mark S. Denning ; Tobias Krämer ; Fernande Grandjean ; Gary J. Long ; Radovan Herchel ; John E. McGrady
Inorganic Chemistry 2010 Volume 49(Issue 13) pp:6160-6171
Publication Date(Web):June 9, 2010
DOI:10.1021/ic100817s
Addition of potassium metal and 2,2,2-crypt (4,7,13,16,21,24-hexaoxa-1,10-diazabicyclo[8.8.8]hexacosane) to a tetrahydrofuran (THF) solution of Fe(2,2′-bipyridine)(mes)2 (1; mes = 2,4,6-Me3C6H2) yielded the anionic complex [Fe(2,2′-bipyridine)(mes)2]− which was isolated as [K(2,2,2-crypt)][Fe(2,2′-bipyridine)(mes)2] (2) alongside the side-product [K(2,2,2-crypt)][Fe(mes)3]·C6H12 (3). A compositionally pure sample of 2 was obtained by dissolving a mixture of 2 and 3 in dry pyridine and layering the resulting solution with toluene. Solid state magnetic susceptibility measurements on 1 reveal Curie−Weiss paramagnetic behavior with a molar magnetic moment of 5.12(1) μB between 20 and 300 K, a value which is in line with the expected iron(II) spin-only value of 4.90 μB. The magnetic measurements carried out on 2 reveal more complex temperature dependent behavior consistent with intramolecular antiferromagnetic coupling (J = −46 cm−1) between the unpaired electrons of the iron(II) ion (SFe = 2) and a π* orbital of the bipyridyl radical (Sbipy = 1/2). Structural data, Mössbauer and electron paramagnetic resonance (EPR) spectroscopic measurements, and density functional theory (DFT) calculations are all consistent with this model of the electronic structure. To the best of our knowledge, species 2 represents the first crystallographically characterized transition metal complex of the 2,2′-bipyridyl ligand for which magnetic, spectroscopic, and computational data indicate the presence of an unpaired electron in the π* antibonding orbital.
Co-reporter:Binbin Zhou ;Dr. Jose M. Goicoechea
Chemistry - A European Journal 2010 Volume 16( Issue 36) pp:11145-11150
Publication Date(Web):
DOI:10.1002/chem.201000507

Abstract

Reaction of cyclooctatetraene (COT) iron(II) tricarbonyl, [Fe(cot)(CO)3], with one equivalent of K4Ge9 in ethylenediamine (en) yielded the cluster anion [Ge8Fe(CO)3]3− which was crystallographically-characterized as a [K(2,2,2-crypt)]+ salt in [K(2,2,2-crypt)]3[Ge8Fe(CO)3]. The chemically-reduced organometallic species [Fe(η3-C8H8)(CO)3] was also isolated as a side-product from this reaction as [K(2,2,2-crypt)][Fe(η3-C8H8)(CO)3]. Both species were further characterized by EPR and IR spectroscopy and electrospray mass spectrometry. The [Ge8Fe(CO)3]3− cluster anion represents an unprecedented functionalized germanium Zintl anion in which the nine-atom precursor cluster has lost a vertex, which has been replaced by a transition-metal moiety.

Co-reporter:Binbin Zhou, Mark S. Denning, Thomas A. D. Chapman, John E. McGrady and Jose M. Goicoechea  
Chemical Communications 2009 (Issue 46) pp:7221-7223
Publication Date(Web):22 Oct 2009
DOI:10.1039/B917185A
Reaction of an ethylenediamine solution of the intermetallic Zintl phase K4Pb9 with dimesitylcadmium yielded the dimeric Zintl ion [Pb9Cd–CdPb9]6− in low crystalline yield. The cluster anion exhibits a Cd–Cd bond and was structurally characterised by single crystal X-ray diffraction in [K(2,2,2-crypt)]6[Cd2Pb18]·2en (1). DFT was also used to explore the bonding in the hexa-anionic cluster.
Co-reporter:Edward Gore-Randall ; Mark Irwin ; Mark S. Denning
Inorganic Chemistry 2009 Volume 48(Issue 17) pp:8304-8316
Publication Date(Web):August 12, 2009
DOI:10.1021/ic9009459
The reaction of ethylenediamine (en) solutions of 2,2′- and 2,4′-bipyridine (bipy) with varying stoichiometric amounts of potassium and rubidium metal resulted in the isolation of compositionally pure solids containing the respective bipyridyl radical anions (2,2′- and 2,4′-bipy•−) and dianions (2,2′- and 2,4′-bipy2−). These species were structurally characterized by single-crystal X-ray diffraction in K(2,2′-bipy)(en) (1a), K4(2,2′-bipy)4(en)4 (1b), Rb2(2,2′-bipy)(en)2 (2), K(2,4′-bipy)(en) (3), K4(2,4′-bipy)2(en)3.5 (4), and Rb4(2,4′-bipy)2(en)3.5 (5). The crystallographic results obtained allow for interesting relationships to be drawn between the electronic structure of the bipyridyl moieties and metric structural data. Further characterization of the solids by means of powder X-ray diffraction, elemental analysis, electron paramagnetic resonance, and IR and Raman spectroscopy is also reported. These studies provide a comprehensive overview of the structural and spectroscopic properties of these often-cited yet elusive air- and moisture-sensitive species, helping to complement the existing data in the chemical literature.
Co-reporter:Binbin Zhou ; Mark S. Denning ; Thomas A. D. Chapman
Inorganic Chemistry 2009 Volume 48(Issue 7) pp:2899-2907
Publication Date(Web):March 5, 2009
DOI:10.1021/ic8022826
Reaction of ethylenediamine solutions of K4E9 (E = Sn, Pb) with diphenylcadmium yielded the Cd(C6H5)-functionalized Zintl ions, closo-[E9Cd(C6H5)]3− (E = Sn (1); Pb (2)). Solution reactivity studies of 1 with tributyltinhydride in pyridine revealed that the cluster is capable of undergoing a coupling reaction to yield the tributyltin-functionalized cluster, closo-{Sn9CdSn[(CH2)3CH3]3}3− (3). In-situ monitoring of this reaction by 1H and 13C{1H} NMR spectroscopy reveals the formation of the tributyltin functionalized cluster anion in addition to free benzene. Species 1−3 were characterized in the solid-state as [K(2,2,2-crypt)]+ salts by single crystal X-ray diffraction and elemental analysis, while the presence of all three cluster anions in solution was confirmed by 1H and 13C{1H} NMR spectroscopy and electrospray mass-spectrometry.
Co-reporter:Mark S. Denning ; ; Mark Irwin ; ;
Inorganic Chemistry 2008 Volume 47(Issue 14) pp:6118-6120
Publication Date(Web):June 17, 2008
DOI:10.1021/ic800726p
The reaction of ethylenediamine solutions of 4,4′-bipyridine with varying stoichiometric amounts of sodium resulted in the isolation of the 4,4′-bipyridyl radical anion (44bipy •−) and the unprecedented 4,4′-bipyridyl dianion (44bipy 2−). The radical was characterized by single-crystal X-ray diffraction in Na(44bipy)(en) ( 1) and Na 2(44bipy) 2(en) 2 ( 2) and the dianion in Na 2(44bipy)(en) 2 ( 3), allowing for interesting correlations to be drawn between electronic structure and metric structural data. Further characterization of the solids by means of powder X-ray diffraction, electron paramagnetic resonance, and/or elemental analysis is also reported.
Co-reporter:Mark S. Denning and Jose M. Goicoechea  
Dalton Transactions 2008 (Issue 43) pp:5882-5885
Publication Date(Web):22 Sep 2008
DOI:10.1039/B810990G
Reaction of an ethylenediamine solution of K4Ge9 with Hg(C6H5)2 yielded [Hg3(Ge9)4]10−, an unprecedented mercury-linked molecular rod consisting of four nine-atom germanium clusters.
Co-reporter:E. N. Faria, A. R. Jupp and J. M. Goicoechea
Chemical Communications 2017 - vol. 53(Issue 52) pp:NaN7095-7095
Publication Date(Web):2017/06/08
DOI:10.1039/C7CC01285C
We describe the reactivity of the 2-phosphaethynolate anion (PCO−) towards enantiomerically pure α-amino acids (AAs) resulting in the formation of novel salts of phosphinecarboxamides bearing chiral functionalities. These transformations occurred quantitatively with all but one of the amino acids trialled (the basic amino acid arginine was found to be unreactive). The resulting ionic species can be readily protonated to afford N-(phosphanyl)carbonyl-amino acids, a novel group of amino acids bearing primary phosphine functionalities.
Co-reporter:M. B. Geeson, A. R. Jupp, J. E. McGrady and J. M. Goicoechea
Chemical Communications 2014 - vol. 50(Issue 82) pp:NaN12284-12284
Publication Date(Web):2014/08/27
DOI:10.1039/C4CC06094F
We describe the coordination chemistry of the primary phosphine PH2C(O)NH2 (phosphinecarboxamide) towards group 6 transition-metals. Experimental and theoretical studies reveal that this novel species has comparable electronic properties to PH3.
Co-reporter:Jordan B. Waters, Thomas A. Everitt, William K. Myers and Jose M. Goicoechea
Chemical Science (2010-Present) 2016 - vol. 7(Issue 12) pp:NaN6987-6987
Publication Date(Web):2016/07/20
DOI:10.1039/C6SC02343F
The room temperature reaction of a 1:1 mixture of phosphorus tribromide (PBr3) and the N-heterocyclic carbene 1,3-bis(2,6-diisopropylphenyl)-imidazol-2-ylidene (IPr) quantitatively affords the Lewis acid–base adduct (IPr)PBr3 (1). Interestingly, when 1 is heated between 55 and 65 °C for a period of several days, dark red crystals slowly begin to form in the reaction vessel accompanied by the release of bromine. The resulting crystalline sample, [P2(IPr)2Br3]Br ([2]Br), results from the reductive coupling of two equivalents of 1, and contains a cationic moiety with a P–P bond that is bridged by a bromine atom. Anion exchange reactions with Na[BArF4] (BArF4 = B(3,5-{CF3}2C6H3)4) afford [2][BArF4]. Abstraction of two equivalents of bromine allows for the isolation of the unprecedented dicationic species [P2(IPr)2Br2]2+ (3) which was isolated and structurally authenticated as two different [BArF4]− salts. Reaction of 2 with mild reductants such as SnBr2 or tetrakis(dimethylamino)ethylene (TDAE) affords [P2(IPr)2Br]+ (4) and the known radical cation [P2(IPr)2]˙+ (5), respectively. These studies show that relatively weak P–Br bonds present in compounds 1–4 can be cleaved in a straightforward manner to afford low oxidation state compounds in high yields.
Co-reporter:Mark S. Denning and Jose M. Goicoechea
Dalton Transactions 2008(Issue 43) pp:NaN5885-5885
Publication Date(Web):2008/09/22
DOI:10.1039/B810990G
Reaction of an ethylenediamine solution of K4Ge9 with Hg(C6H5)2 yielded [Hg3(Ge9)4]10−, an unprecedented mercury-linked molecular rod consisting of four nine-atom germanium clusters.
Co-reporter:Robert S. P. Turbervill and Jose M. Goicoechea
Chemical Communications 2012 - vol. 48(Issue 49) pp:NaN6102-6102
Publication Date(Web):2012/04/23
DOI:10.1039/C2CC32244G
Dimethylformamide solutions of K3E7 (E = P, As) react with acetylene yielding the 1,2,3-tripnictolide anions [E3C2H2]− (R = P (1), As (2)). Preliminary studies have shown that 1 and 2 displace labile ligands in [Ru(COD){η3-CH3C(CH2)2}2] (COD = 1,5-cyclooctadiene) to yield the novel complexes [Ru(η5-E3C2H2){CH3C(CH2)2}2}]− (E = P (3), As (4)).
Co-reporter:Caroline M. Knapp, B. H. Westcott, M. A. C. Raybould, J. E. McGrady and J. M. Goicoechea
Chemical Communications 2012 - vol. 48(Issue 100) pp:NaN12185-12185
Publication Date(Web):2012/11/01
DOI:10.1039/C2CC36888A
Reaction of an ethylenediamine solution of K3As7 with the low-valent, low-coordinate cobalt(II) complex [Co(mes)2(PEt2Ph)2] yielded the novel dianionic species [Co(η3-As3){η4-As4(mes)2}]2− (1). The [η4-As4(mes)2]2− moiety present in 1 is a rare example of a group 15 analogue of a butadienediide.
Co-reporter:Caroline M. Knapp, Joseph S. Large, Nicholas H. Rees and Jose M. Goicoechea
Chemical Communications 2011 - vol. 47(Issue 14) pp:NaN4113-4113
Publication Date(Web):2011/03/07
DOI:10.1039/C1CC10236B
Reaction of K3P7, FeCl2, [NH4][B(C6H5)4] and 2,2,2-crypt yielded the unprecedented cluster dianion [Fe(HP7)2]2−. This species was characterized by single crystal X-ray diffraction, multielement NMR spectroscopy and electrospray mass-spectrometry. DFT calculations on the dianion were also conducted.
Co-reporter:Binbin Zhou, Mark S. Denning, Thomas A. D. Chapman, John E. McGrady and Jose M. Goicoechea
Chemical Communications 2009(Issue 46) pp:NaN7223-7223
Publication Date(Web):2009/10/22
DOI:10.1039/B917185A
Reaction of an ethylenediamine solution of the intermetallic Zintl phase K4Pb9 with dimesitylcadmium yielded the dimeric Zintl ion [Pb9Cd–CdPb9]6− in low crystalline yield. The cluster anion exhibits a Cd–Cd bond and was structurally characterised by single crystal X-ray diffraction in [K(2,2,2-crypt)]6[Cd2Pb18]·2en (1). DFT was also used to explore the bonding in the hexa-anionic cluster.
Co-reporter:Adinarayana Doddi, Michael Weinhart, Alexander Hinz, Dirk Bockfeld, Jose M. Goicoechea, Manfred Scheer and Matthias Tamm
Chemical Communications 2017 - vol. 53(Issue 45) pp:NaN6072-6072
Publication Date(Web):2017/04/25
DOI:10.1039/C7CC02628E
N-Heterocyclic carbene adducts of the parent arsinidene (AsH) were prepared by two different synthetic routes, either by reaction of As(SiMe3)3 with 2,2-difluoroimidazolines followed by desilylation or by reaction of [Na(dioxane)3.31][AsCO] with imidazolium chlorides.
Co-reporter:Jordan B. Waters and Jose M. Goicoechea
Dalton Transactions 2014 - vol. 43(Issue 38) pp:NaN14248-14248
Publication Date(Web):2014/05/06
DOI:10.1039/C4DT00954A
Reaction of the lithiated N-heterocyclic carbene [:C[N(2,6-iPr2C6H3)]2(CH)CLi]n (LiIPr) with KOtBu in diethylether (Et2O) afforded the novel organo-potassium compound [:C[N(2,6-iPr2C6H3)]2(CH)CK(THF)2] (KIPr·2THF). Both LiIPr and KIPr can be interpreted as ditopic carbanionic carbenes (or alkali metal salts of anionic “dicarbenes”) and are interesting precursors for the synthesis of novel metal complexes bearing carbanionic carbenes as ligands. Reaction of KIPr with M[N(SiMe3)2]2 (M = Zn, Sn) afforded salts of the anionic three coordinate complexes [M{C(CH)[N(2,6-iPr2C6H3)]2C:}{N(SiMe3)2}2]− (M = Zn (1) and Sn (2)). Contrasting reactivity was observed for the other group 14 bis-amide compounds M[N(SiMe3)2]2 (M = Ge, Pb), which initially appear to yield analogous 1:1 complexes (M = Ge (3) and Pb (4)), however over time give rise to compounds bearing two ditopic carbanionic carbenes ([M{C(CH)[N(2,6-iPr2C6H3)]2C:}2{N(SiMe3)2}]−; Ge (5) and Pb (6)) and the tris-amide anions ([M{N(SiMe3)3}]−), presumably via a Schlenk-type equilibrium. Compounds 5 and 6 can be directly synthesized by reacting M[N(SiMe3)2]2 (M = Ge, Pb) with two equivalents of KIPr, respectively.
Co-reporter:Robert S. P. Turbervill and Jose M. Goicoechea
Chemical Communications 2012 - vol. 48(Issue 10) pp:NaN1472-1472
Publication Date(Web):2011/06/16
DOI:10.1039/C1CC12089A
Reaction of the hydrogenheptaphosphide dianion, [HP7]2−, with one equivalent of bis(2,6-diisopropylphenyl)carbodiimide yielded the exo-functionalized cluster [P7C(NDipp)(NHDipp)]2− (Dipp = 2,6-diisopropylphenyl). This species was characterized by single crystal X-ray diffraction, multielement NMR spectroscopy and electrospray mass-spectrometry. DFT calculations on the dianion were also conducted.
Co-reporter:Gabriela Espinoza Quintero, Isabelle Paterson-Taylor, Nicholas H. Rees and Jose M. Goicoechea
Dalton Transactions 2016 - vol. 45(Issue 5) pp:NaN1936-1936
Publication Date(Web):2015/07/17
DOI:10.1039/C5DT02380G
Reactions of the protonated heptaphosphide dianion, [HP7]2−, with one equivalent of E[N(SiMe3)2]2 (E = Ge, Sn, Pb) give rise to novel derivatized cluster anions [P7EN(SiMe3)2]2− (E = Ge (1), Sn (2) and Pb (3)). All three species were characterized by multi-element solution-phase NMR spectroscopy and electrospray ionization mass spectrometry. In addition, 1 and 2 were structurally authenticated by means of single crystal X-ray diffraction in [K(18-crown-6)]2[P7EN(SiMe3)2]·2py. Interestingly, while 2 appears to be indefinitely stable in solution for prolonged periods of time, the germanium-containing analogue, 1, readily decomposes at room temperature giving rise to the dimeric species [(P7Ge)2N(SiMe3)2]3− (4) and [K(18-crown-6)][N(SiMe3)2]. A low quality single crystal X-ray structure of the former allowed for the confirmation of its composition and connectivity which is consistent with the 31P NMR spectrum obtained for the anion.
Co-reporter:Tobias Krämer, Jack C. A. Duckworth, Matthew D. Ingram, Binbin Zhou, John E. McGrady and Jose M. Goicoechea
Dalton Transactions 2013 - vol. 42(Issue 34) pp:NaN12129-12129
Publication Date(Web):2013/04/08
DOI:10.1039/C3DT50643F
The synthesis of a new endohedral ten-vertex Zintl ion cluster, [Fe@Sn10]3−, isoelectronic with [Fe@Ge10]3−, is reported. In an attempt to place this new cluster within the context of the known structural chemistry of the M@E10 family (M = transition metal, E = main group element), we have carried out a detailed electronic structure analysis of the different structural types: viz bicapped square antiprismatic ([Ni@Pb10]2−, [Zn@In10]8−), tetra-capped trigonal prismatic ([Ni@In10]10−) and the remarkable pentagonal prismatic [Fe@Ge10]3− and [Co@Ge10]3−. We establish that the structural trends can be interpreted in terms of a continuum of effective electron counts at the E10 cage, ranging from electron deficient (<4n + 2) in [Ni@In10]10− to highly electron rich (>4n + 2) in [Fe@Ge10]3−. The effective electron count differs from the total valence electron count in that it factors in the increasingly active role of the metal d electrons towards the left of the transition series. The preference for a pentagonal prismatic geometry in [Fe@Ge10]3− emerges as a natural consequence of backbonding to the cage from four orthogonal 3d orbitals of the low-valent metal ion. Our calculations suggest that the new [Fe@Sn10]3− cluster should also exhibit structural consequences of backbonding from the metal to the cage, albeit to a less extreme degree than in its Ge analogue. The global minimum lies on a very flat surface connecting D4d, C2v and C3v-symmetric minima, suggesting a very plastic structure that may be easily deformed by the surrounding crystal environment. If so, then this provides a new and quite distinct structural type for the M@E10 family.
Co-reporter:Rebecca A. Musgrave, Robert S. P. Turbervill, Mark Irwin, Radovan Herchel and Jose M. Goicoechea
Dalton Transactions 2014 - vol. 43(Issue 11) pp:NaN4344-4344
Publication Date(Web):2013/10/16
DOI:10.1039/C3DT52638K
Reaction of dimesityliron(II) (Fe2(mes)4) with the N-heterocyclic carbenes 1,3-bis(2,6-diisopropylphenyl)-imidazol-2-ylidene (IPr) and 1,3-bis(2,6-dimethylphenyl)hexahydropyrimidin-2-ylidene (6-Xyl) afforded the novel trigonal planar complexes [Fe(IPr)(mes)2] (1) and [Fe(6-Xyl)(mes)2] (2), respectively. Both species were structurally characterized by single crystal X-ray diffraction and display structures and magnetic responses consistent with a quintet ground state (S = 2). Reaction of 1 with KC8 in THF afforded K+ salts of the anionic complex [{:C[N(2,6-iPr2C6H3)]2(CH)C}2Fe(mes)]− (3) and the homoleptic organometallic anion [Fe(mes)3]− (4). By contrast, reduction of 2 resulted in extensive decomposition and intractable product mixtures. Complex 3 is coordinated by two ditopic carbanionic carbenes via the C4/C5 position while the C2 position retains unquenched carbenic character and remains vacant for further coordination. This was corroborated by reacting solutions of 3 with one and two equivalents of triethylaluminium (AlEt3) which resulted in the formation of [{Et3Al:C[N(2,6-iPr2C6H3)]2(CH)C}{:C[N(2,6-iPr2C6H3)]2(CH)C}Fe(mes)]− (5) and [{Et3Al:C[N(2,6-iPr2C6H3)]2(CH)C}2Fe(mes)]− (6), respectively. Both of these species were structurally characterized as [K(2,2,2-crypt)]+ salts.
1,4,7,10-Tetraazacyclododecane, 1,4,7,10-tetrakis(2-pyridinylmethyl)-