Co-reporter:Mustafa A-jabbar Al-jumaili, Simon Woodward
Tetrahedron 2017 Volume 73, Issue 40(Issue 40) pp:
Publication Date(Web):5 October 2017
DOI:10.1016/j.tet.2017.08.026
Derivatives of 3,4-ethylenedithiothiophene (EDTT) are reported starting from tetrabromothiophene. Selective 2,5-dilithiation followed by reaction with a range of aldehydes gives diols as mixtures of diastereomers. Only the 2 and 5 positions in thiophene react leaving the 3,4-bromides for further elaboration. The diols are oxidised to their corresponding diketones using activated MnO2. Reaction with 1,2-ethanedithiol, by addition-elimination, provides access to novel monomers for the preparation of conjugated copolymers of 3,4-ethylenedithiothiophene (EDTT). A range of these monomers can be attained by applying the synthesis of a series of ketones applicable to further synthesis of π-extended thiophene-based organic semiconductors. Finally, this new route was compared to 3,4-ethylenedioxythiophene (EDOT) dialdehyde derivatives synthesised by an alternative to literature chemistry.Download high-res image (97KB)Download full-size image
Co-reporter:Laurence Burroughs, John Ritchie, Simon Woodward
Tetrahedron 2016 Volume 72(Issue 13) pp:1686-1689
Publication Date(Web):31 March 2016
DOI:10.1016/j.tet.2016.02.025
The reaction pathway for the formation of tetracenes from the diols 1,2-C6H4(CHOHCCAr)2, LiHDMS, CS2 and MeI has been modelled by computational methods at the CBS-QB3 level of theory. Comparison of PhCHOC(=S)YCCPh (Y=S− or SMe) indicates a slight kinetic advantage for the anionic system towards [3,3]-sigmatropic rearrangement [Eact(calcd) 19.7 vs 21.8 kcal mol−1]. Using anthracene-based models, 10-{SC(=O)Y}-4a,10-dihydroanthracene (Y=S− or SMe), allows direct comparison of both syn and anti-manifolds in the neutral versus anionic Chugaev elimination. syn-Elimination of [HSC(=O)S]- is distinctly favoured [Eact(calcd) 11.4 kcal mol−1] versus syn elimination of neutral methylated HSC(=O)SMe [Eact(calcd) 27.5 kcal mol−1]. The smaller barrier to syn elimination of the anionic leaving group is in accord with the low temperature conditions required for this Chugaev reaction (60 °C) and suggests a general advantage in carrying out Chugaev eliminations in anionic manifolds.
Co-reporter:Lee Eccleshare;Leticia Lozada-Rodríguez;Phillippa Cooper;Dr. Laurence Burroughs;John Ritchie;Dr. William Lewis ;Dr. Simon Woodward
Chemistry - A European Journal 2016 Volume 22( Issue 35) pp:12542-12547
Publication Date(Web):
DOI:10.1002/chem.201601970
Abstract
Sequential treatment of 2-C6H4Br(CHO) with LiC≡CR1 (R1=SiMe3, tBu), nBuLi, CuBr⋅SMe2 and HC≡CCHClR2 [R2=Ph, 4-CF3Ph, 3-CNPh, 4-(MeO2C)Ph] at −50 °C leads to formation of an intermediate carbanion (Z)-1,2-C6H4{CA(=O)C≡CBR1}{CH=CH(CH−)R2} (4). Low temperatures (−50 °C) favour attack at CB leading to kinetic formation of 6,8-bicycles containing non-classical C-carbanion enolates (5). Higher temperatures (−10 °C to ambient) and electron-deficient R2 favour retro σ-bond C−C cleavage regenerating 4, which subsequently closes on CA providing 6,6-bicyclic alkoxides (6). Computational modelling (CBS-QB3) indicated that both pathways are viable and of similar energies. Reaction of 6 with H+ gave 1,2-dihydronaphthalen-1-ols, or under dehydrating conditions, 2-aryl-1-alkynylnaphthlenes. Enolates 5 react in situ with: H2O, D2O, I2, allylbromide, S2Me2, CO2 and lead to the expected C-E derivatives (E=H, D, I, allyl, SMe, CO2H) in 49–64 % yield directly from intermediate 5. The parents (E=H; R1=SiMe3, tBu; R2=Ph) are versatile starting materials for NaBH4 and Grignard C=O additions, desilylation (when R1=SiMe) and oxime formation. The latter allows formation of 6,9-bicyclics via Beckmann rearrangement. The 6,8-ring iodides are suitable Suzuki precursors for Pd-catalysed C−C coupling (81–87 %), whereas the carboxylic acids readily form amides under T3P® conditions (71–95 %).
Co-reporter:Christopher M. Latham;William Lewis;Alexer J. Blake
European Journal of Organic Chemistry 2015 Volume 2015( Issue 8) pp:1819-1823
Publication Date(Web):
DOI:10.1002/ejoc.201403560
Abstract
The amidoamines (S)-Ar1CONHCHRCH2NHAr2 [Ar1 = o-C6H4SO3H, R = Bn, iBu, iPr; Ar2 = 2,6-iPr2C6H4, 2,6-Et2C6H4, 2,4,6-Me3C6H2] cyclise to (S) 1-aryl-substituted 4,5-dihydro-1H-imidazolinium species with HC(OEt)3 in moderate-to-excellent yields on heating to 150–175 °C (nine examples, four isolated yields of 48 to >97 %). They are attained as their o-C6H4(SO3–)(CO2Et) salts. The latter are readily deprotonated to afford analytically pure (S) 1-aryl-substituted 4,5-dihydro-1H-imidazoles (imidazolines). The purification of the intermediate sulfonate salts is not always necessary, and analytically pure imidazolines are isolated by simple kugelrohr distillation (nine examples, 45–95 %) after basification. Imidazoline alkylation provides a library of (S)-N-alkylimidazolinium salts (23 examples, 74–97 %). As the initially required amidoamines are available in simple one-pot reactions, the overall approach constitutes a rather efficient approach to this useful family of chiral N-heterocyclic carbene (NHC) ligand precursors (effectively three steps from commercial N-Boc-α-amino alcohols; BOC = tert-butyloxycarbonyl).
Co-reporter:Benoit Wahl;Darren S. Lee
European Journal of Organic Chemistry 2015 Volume 2015( Issue 27) pp:6033-6039
Publication Date(Web):
DOI:10.1002/ejoc.201500730
Abstract
The diesters (E)-RO2CCH=CHCO2R [R = Et, CH2Ph, iPr, iBu, tBu, 2-ethylhexyl, (–)-menthyl, (–)-bornyl] undergo 1,4-trichloromethylation with Me3SiCCl3 in 84–95 % yield with Bu4NX (X = OAc, Cl, F; 0.5–10 mol-%) as an initiator. Optimal catalysis is attained with X = OAc and F; chloride is less effective, bromide is inadequate. Poor regioselectivity but high yields are demonstrated by mixed diesters (2 examples, both > 80 %). Heating of the trichloromethyl adducts, in the presence of Bu4N(OAc), leads to equilibrating, approximately equimolar, mixtures of RO2CC(=CCl2)CHCO2R and RO2CCH=CClCHClCO2R above 100 °C (7 examples). The latter component is produced under thermodynamic control, which can be isolated pure in some cases. The former is the kinetic product of E2 elimination.
Co-reporter:Darren S. Lee, Zacharias Amara, Martyn Poliakoff, Thomas Harman, Gary Reid, Barrie Rhodes, Steve Brough, Thomas McInally, and Simon Woodward
Organic Process Research & Development 2015 Volume 19(Issue 7) pp:831-840
Publication Date(Web):June 24, 2015
DOI:10.1021/acs.oprd.5b00101
Methods for the batch scale up of DABAL-Me3 promoted direct ester to amide synthesis have been demonstrated at 10–100 g scales using a tert-amide model compound. Procedures for 20 g scale couplings in standard laboratory glassware and up to 0.1 kg in industry-standard jacketed glass reactors in near quantitative yields are given. A derivative of the anticancer agent Imatinib (Gleevec) has been synthesized on a 26 g scale (98% yield, >98% purity) establishing DABAL-Me3 as a potential alternative for the synthesis of amides in API scale preparations. Continuous flow methodology provides a method for larger scales (productivities of >50 g h–1). In addition, nitriles were coupled to primary amines and hydrazines with DABAL-Me3, resulting in the clean formation of free amidines (16 examples) and amidrazones.
Co-reporter:Dr. Laurence Burroughs;Lee Eccleshare;John Ritchie;Omkar Kulkarni;Dr. Barry Lygo;Dr. Simon Woodward;Dr. William Lewis
Angewandte Chemie International Edition 2015 Volume 54( Issue 36) pp:10648-10651
Publication Date(Web):
DOI:10.1002/anie.201505347
Abstract
An intramolecular Cannizzaro-type hydride transfer to an in situ prepared allene enables the synthesis of ortho-fused 4-substituted cycloocta-2,5-dien-1-ones with unprecedented technical ease for an eight-ring carboannulation. Various derivatives could be obtained from commercially available (hetero)aryl aldehydes, trimethylsilylacetylene, and simple propargyl chlorides in good yields.
Co-reporter:Dr. Laurence Burroughs;Lee Eccleshare;John Ritchie;Omkar Kulkarni;Dr. Barry Lygo;Dr. Simon Woodward;Dr. William Lewis
Angewandte Chemie 2015 Volume 127( Issue 36) pp:10794-10797
Publication Date(Web):
DOI:10.1002/ange.201505347
Abstract
An intramolecular Cannizzaro-type hydride transfer to an in situ prepared allene enables the synthesis of ortho-fused 4-substituted cycloocta-2,5-dien-1-ones with unprecedented technical ease for an eight-ring carboannulation. Various derivatives could be obtained from commercially available (hetero)aryl aldehydes, trimethylsilylacetylene, and simple propargyl chlorides in good yields.
Co-reporter:D. Willcox, S. Woodward and A. Alexakis
Chemical Communications 2014 vol. 50(Issue 14) pp:1655-1657
Publication Date(Web):07 Jan 2014
DOI:10.1039/C3CC48191C
Chloromethylvinyl alanes (E)-ClMeAl(CHCHR) prepared directly from terminal alkynes undergo 1,4-addition to cyclohexenone and 3-methylcyclohexenone in moderate to good yield (30–70%) and good to excellent stereoselectivity (80–98% ee) using readily available copper(I) sources and chiral ligands.
Co-reporter:Benoit Wahl, Albert Cabré, Simon Woodward, William Lewis
Tetrahedron Letters 2014 Volume 55(Issue 42) pp:5829-5831
Publication Date(Web):15 October 2014
DOI:10.1016/j.tetlet.2014.08.122
Nucleophilic addition of readily available TMSCCl3 to N-phosphinoyl benzaldimines allows preparation of N-phosphinoyl-α-(trichloromethyl)benzylamines. Typically, the reaction in THF at room temperature using tetrabutylammonium difluorotriphenylsilicate (TBAT) as a catalytic promoter, afforded very good yields (65–95% range) for most derivatives within 1 h at room temperature.
Co-reporter:Darren Willcox, Humaira Gondal, Marc Garcia Civit, Simon Woodward
Tetrahedron Letters 2014 Volume 55(Issue 10) pp:1720-1721
Publication Date(Web):5 March 2014
DOI:10.1016/j.tetlet.2014.01.105
Rare alkenylalanes are prepared by Cp2TiCl2 or Cp*2ZrCl2 (Cp = η-C5H5; Cp* = η-C5Me5) catalysed addition of HAlCl2·(THF)2 to terminal alkynes (R1CCH; R1 = alkyl). Use of minimum head-volume sealed vials maximises the hydroalumination yields of volatile alkynes. Facile 1,4-addition of the resultant alkenylalanes to unsaturated malonates R2CHC(CO2R3)2 (R2 = alkyl, aryl, R3 = alkyl) is observed providing rapid and convenient access to the addition products.
Co-reporter:Dr. Na Wu;Dr. Benoit Wahl;Dr. Simon Woodward;Dr. William Lewis
Chemistry - A European Journal 2014 Volume 20( Issue 25) pp:7718-7724
Publication Date(Web):
DOI:10.1002/chem.201402394
Abstract
Improved synthetic conditions allow preparation of TMSCCl3 in good yield (70 %) and excellent purity. Compounds of the type NBu4X [X=Ph3SiF2 (TBAT), F (tetrabutylammonium fluoride, TBAF), OAc, Cl and Br] act as catalytic promoters for 1,4-additions to a range of cyclic and acyclic nitroalkenes, in THF at 0–25 °C, typically in moderate to excellent yields (37–95 %). TBAT is the most effective promoter and bromide the least effective. Multinuclear NMR studies (1H, 19F, 13C and 29Si) under anaerobic conditions indicate that addition of TMSCCl3 to TBAT (both 0.13 M) at −20 °C, in the absence of nitroalkene, leads immediately to mixtures of Me3SiF, Ph3SiF and NBu4CCl3. The latter is stable to at least 0 °C and does not add nitroalkene from −20 to 0 °C, even after extended periods. Nitroalkene, in the presence of TMSCCl3 (both 0.13 M at −20 °C), when treated with TBAT, leads to immediate formation of the 1,4-addition product, suggesting the reaction proceeds via a transient [Me3Si(alkene)CCl3] species, in which (alkene) indicates an Si⋅⋅⋅O coordinated nitroalkene. The anaerobic catalytic chain is propagated through the kinetic nitronate anion resulting from 1,4 CCl3− addition to the nitroalkene. This is demonstrated by the fact that isolated NBu4[CH2NO2] is an efficient promoter. Use of H2CCH(CH2)2CHCHNO2 in air affords radical-derived bicyclic products arising from aerobic oxidation.
Co-reporter:Philip Andrews, Christopher M. Latham, Marc Magre, Darren Willcox and Simon Woodward
Chemical Communications 2013 vol. 49(Issue 15) pp:1488-1490
Publication Date(Web):14 Jan 2013
DOI:10.1039/C2CC37537K
Stabilized AlHCl2·(THF)2 hydroaluminates RCCH with exceptional chemo-, regio- and stereoselectivity under efficient ZrCl2(η-C5Me5)2 catalysis (2–5 mol%). The resulting vinyl alanes undergo palladium cross-coupling with a wide range of sp2 electrophiles (aryl, heteroaryl and vinyl halides/pseudohalides) in good to excellent yields.
Co-reporter:Melchior Cini;Tracey D. Bradshaw;William Lewis
European Journal of Organic Chemistry 2013 Volume 2013( Issue 19) pp:3997-4007
Publication Date(Web):
DOI:10.1002/ejoc.201300474
Abstract
The copper-catalysed (10 mol-% CuBr·SMe2, CuCN·LiCl or CuI/PPh3) addition of RMgBr to the pentafulvene 1-(cyclopenta-2,4-dien-1-ylidenemethyl)-2-methoxybenzene allows the formation of cyclopentadienyl derivatives with α-CHR(2-MeOPh) sidechains (R = Me, Et, nBu, iBu, allyl, Ph) without H– transfer. The deprotonation of these sec-alkyl-substituted cyclopentadienyls followed by the addition of TiCl4 allows the isolation of TiCl2{η5-C5H4CHR(2-OMePh)} as rac/meso mixtures that show activity against human colon, breast and pancreatic cell lines (GI50 2.3–42.4 μM).
Co-reporter:Nathalie Dubois, Daniel Glynn, Thomas McInally, Barrie Rhodes, Simon Woodward, Derek J. Irvine, Chris Dodds
Tetrahedron 2013 69(46) pp: 9890-9897
Publication Date(Web):
DOI:10.1016/j.tet.2013.08.062
Co-reporter:Rosemary H. Crampton, Martin Fox, Simon Woodward
Tetrahedron: Asymmetry 2013 Volume 24(9–10) pp:599-605
Publication Date(Web):31 May 2013
DOI:10.1016/j.tetasy.2013.04.006
The sequential reaction of chlorosulfonyl isocyanate with t-BuOH, t-BuNH2 and TFA allows formation of H2NSO2NHBut. Condensation of the latter with Ar1CHO in the presence of Ti(OEt)4 provides the activated imines Ar1CHNSO2NHBut (59–89%). Commercially available boronic acids add to these imines with good stereoselectivity (76–98% ee) using readily available diene ligands. Simple deprotection with 5% w/w water in pyridine affords free Ar1CHNH2Ar2.(S)-N-tert-Butyl-N′-(4-chlorophenyl)(phenyl)methylsulfamideC17H21ClN2O2See = 81%[α]D=-1.9[α]D=-1.9 (c 1.09, CHCl3)Source of chirality: Asymmetric catalysisAbsolute configuration: (1S)(R)-N-tert-Butyl-N′-(4-fluorophenyl)(phenyl)methylsulfamideC17H21FN2O2See = 95%[α]D=+0.5[α]D=+0.5 (c 0.97, CHCl3)Source of chirality: Asymmetric catalysisAbsolute configuration: (1R)(R)-N-tert-Butyl-N′-(R)-(4-trifluoromethylphenyl)(phenyl)methylsulfamideC18H21F3N2O2See = 98%[α]D=-3.7[α]D=-3.7 (c 1.35, CHCl3)Source of chirality: Asymmetric catalysisAbsolute configuration: (1R)(R)-N-tert-Butyl-N′-(4-methylphenyl)(phenyl)methylsulfamideC18H24N2O2See = 95%[α]D=+3.9[α]D=+3.9 (c 0.99, CHCl3)Source of chirality: Asymmetric catalysisAbsolute configuration: (1R)(S)-N-tert-Butyl-N′-(3-methylphenyl)(phenyl)methylsulfamideC18H24N2O2See = 85%[α]D=-2.0[α]D=-2.0 (c 1.10, CHCl3)Source of chirality: Asymmetric catalysisAbsolute configuration: (1S)(R)-N-tert-Butyl-N′-(3-methoxyphenyl)(phenyl)methylsulfamideC18H24N2O3See = 97%[α]D=-1.0[α]D=-1.0 (c 1.26, CHCl3)Source of chirality: Asymmetric catalysisAbsolute configuration: (1R)
Co-reporter:Christopher M. Latham;Alexer J. Blake;William Lewis;Matthew Lawrence
European Journal of Organic Chemistry 2012 Volume 2012( Issue 4) pp:699-707
Publication Date(Web):
DOI:10.1002/ejoc.201101336
Abstract
A one-pot reaction of Boc-protected amino alcohols and 2-sulfobenzoic anhydride followed by the addition of a wide variety of primary amines has allowed rapid access to diverse libraries of amidosulfonates 1,2-C6H4(SO3–)(COCHR1CH2NH2+R2) (R1 = Bn, iPr, Ph; R2 = aryl, CHMePh, 1-adamantyl), of which two examples have been characterised by X-ray crystallography. These reactions proceed by SN2 opening of unisolated oxazoline intermediates. The amidosulfonates have been converted to imidazolinium sulfonate zwitterions (N-heterocyclic carbene precursors) in a two-step process, which involved reduction with BH3·SMe2 and cyclisation with HC(OEt)3. Two examples (R1 = Bn, R2 = 2,6-iPrC6H3 and R1 = iPr, R2 = 3,5-Me2C6H3) have been crystallographically characterised, the latter as a PF6 salt. The imidazolinium sulfonate zwitterions (1 mol-%) catalysed additions of RMgBr (R = alkyl) to (E)-ArC(R)=CHCH2Br (R = H, Me). Additions to cinnamyl bromides showed high γ (SN2′) selectivity (up to 18.4:1) and provided significantly enantiomerically enriched products (up to 82 % ee).
Co-reporter:Rosemary Crampton;Martin Fox
Advanced Synthesis & Catalysis 2011 Volume 353( Issue 6) pp:903-906
Publication Date(Web):
DOI:10.1002/adsc.201000838
Abstract
Bis-sulfamyl imines are shown to be potentially ideal substrates for rhodium-catalysed asymmetric additions of arylboron nucleophiles as they show: (i) near perfect enantioselectivities (11 examples, 98–99+% ee), (ii) good to excellent diastereoselectivities (10–32:1 rac:meso), and (iii) high functional group tolerance in removal of the low molecular weight protecting group via mild heating in aqueous pyridine.
Co-reporter:Simon Woodward, Montserrat Diéguez, Oscar Pàmies
Coordination Chemistry Reviews 2010 Volume 254(17–18) pp:2007-2030
Publication Date(Web):September 2010
DOI:10.1016/j.ccr.2010.03.005
This review describes recent development in the use of sugar-derived ligands in the selective synthesis of organic molecules. Developments in the recent literature (2004–2009) are highlighted in the areas of hydrogenation, 1,2- and 1,4-additions of nucleophiles to CO and CNR based substrates, cross-coupling, hydroformylation, oxidation and other reactions. Connections to earlier studies are also noted were relevant. Some suggestions as to the underlying features that make sugar-based ligands highly useful modular ligands in selective catalysis are given. Finally, advice is presented (for the non-specialist) on optimal entry points and basic starting materials for sugar-ligand synthesis.
Co-reporter:Matthias Welker, Simon Woodward and Alexandre Alexakis
Organic Letters 2010 Volume 12(Issue 3) pp:576-579
Publication Date(Web):January 8, 2010
DOI:10.1021/ol9027682
An opening of vinyl oxiranes has been accomplished with Zn and Al enolates resulting from asymmetric conjugate addition reactions on cyclic enones. This novel tandem procedure affords the adducts in moderate to good yields, enantioselectivities up to 98%, and moderate to good cis/trans selectivities. This provides potentially useful synthetic substrates to prepare complex bicyclic compounds.
Co-reporter:Martta Asikainen, William Lewis, Alexander J. Blake, Simon Woodward
Tetrahedron Letters 2010 Volume 51(Issue 49) pp:6454-6456
Publication Date(Web):8 December 2010
DOI:10.1016/j.tetlet.2010.10.013
Reaction of cuprates derived from R3MgBr/CuI/LiBr (R3 = n-alkyl) with R1CCCH(O2CR2)2 (R1 = sp2 hybridised substituent, R2 = mainly Me, alkyl, Ph) provides access to allenyl esters R1R3CCCH(O2CR2) (51–88%). Such species are not accessible via rearrangement of precursor propargylic R1R3C(O2CR2)CCH.The otherwise unknown allenic acetate motif ArRCCCH(OAc) can be accessed by an SN2′ acetate displacement from RCCCH(OAc)2 by RMgBr/CuI/LiBr derived cuprates.
Co-reporter:Dr. Christine Hawner;Daniel Müller;Ludovic Gremaud;Abdellah Felouat; Simon Woodward; Alexre Alexakis
Angewandte Chemie International Edition 2010 Volume 49( Issue 42) pp:7769-7772
Publication Date(Web):
DOI:10.1002/anie.201003300
Co-reporter:Matthias Welker, Simon Woodward
Tetrahedron 2010 66(52) pp: 9954-9963
Publication Date(Web):
DOI:10.1016/j.tet.2010.10.048
Co-reporter:Daniel Glynn;Jonathan Shannon Dr.
Chemistry - A European Journal 2010 Volume 16( Issue 3) pp:1053-1060
Publication Date(Web):
DOI:10.1002/chem.200901803
Abstract
A practical asymmetric 1,2-addition of functionalised arylzinc halides to aromatic and aliphatic aldehydes is described by the use of aminoalcohol catalysis in the presence of AlMe3. The process is simple to carry out, uses only commercially available reagents/ligands and provides moderate to good (80–96 % ee) enantioselectivities for a wide range of substrates. Either commercial ArZnX reagents or those prepared in situ from low cost aryl bromides can be used. In the latter case electrophilic functional groups are tolerated (CO2Et, CN). The reaction relies on rapid exchange between ArZnX and AlMe3 to generate mixed organometallic species that lead to the formation of a key intermediate that is distinctly different from the classic “anti” transition states of Noyori. NMR monitoring and related experiments have been used to probe the validity of the proposed selective transition state.
Co-reporter:Matthias Welker ;LuisF. Veiros ;MariaJosé Calhorda
Chemistry - A European Journal 2010 Volume 16( Issue 19) pp:5620-5629
Publication Date(Web):
DOI:10.1002/chem.200903310
Abstract
Reaction of the cyclic thioacetal (RS)2CHCHO [R=1/2×(CH2)3] with HCCCOMe, followed by treatment with TsCl/DABCO (Ts=tosyl, DABCO=1,4-diazabicyclo[2.2.2]octane) affords the mono-protected 1,4-benzoquinone dithioacetal. The reactivity of this SR-protected 1,4-benzoquinone has been compared with the behavior of the analogous OR-protected acetal in copper-catalyzed additions of ZnMe2 by using chiral phosphoramidite ligands. The activation energy for 1,4-methylation of the latter OR-acetals with ZnMe2 (>95 % ee) has been determined for two CuX2 pre-catalysts (X=OAc, 12.2 kcal mol−1; X=OTf, 6.7 kcal mol−1; Tf=triflate). The dithioacetal SR aromatizes in the presence of CuI/ZnMe2 giving 1,4-HOC6H4S(CH2)3SMe through CS bond formation. The disparate behavior of these two very closely related substrates is in accordance with the formation of closely related cuprate intermediates that were optimized by DFT calculations, supporting the synthetic and kinetic studies and thus defining the mechanisms of both pathways.
Co-reporter:Dr. Christine Hawner;Daniel Müller;Ludovic Gremaud;Abdellah Felouat; Simon Woodward; Alexre Alexakis
Angewandte Chemie 2010 Volume 122( Issue 42) pp:7935-7938
Publication Date(Web):
DOI:10.1002/ange.201003300
Co-reporter:Xiaoping Tang;Norbert Krause
European Journal of Organic Chemistry 2009 Volume 2009( Issue 17) pp:2836-2844
Publication Date(Web):
DOI:10.1002/ejoc.200900226
Abstract
A new catalytic method for the synthesis of α-hydroxyallenes is described. Efficient SN2′ substitution of propargylic dioxolanones has been achieved with a copper(I)/P(OBu)3 catalyst using Grignard reagents as the nucleophiles. The reaction tolerates a wide variety of propargylic dioxolanones, the corresponding primary and secondary α-hydroxyallenes are obtained in good to excellent yields and excellent diastereoselectivity. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2009)
Co-reporter:Andrew Novak;Maria José Calhorda;Paulo Jorge Costa
European Journal of Organic Chemistry 2009 Volume 2009( Issue 6) pp:898-903
Publication Date(Web):
DOI:10.1002/ejoc.200801029
Abstract
Enantioselective Ni-catalysed methylation of Baylis–Hillman-derived allylic electrophiles in the presence of ferrophite ligands has been investigated computationally and experimentally. The sense and degree of enantioselectivity attained is independent of both the leaving group and the isomeric structure of the initial allylic halide. DFT studies support the selective formation of a limited number of energetically favoured anti and syn π-allyl intermediates. The observed regio- and enantioselectivity can be rationalised based on the energetics of these structures. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2009)
Co-reporter:Rosemary H. Crampton, Samir El Hajjaji, Martin E. Fox, Simon Woodward
Tetrahedron: Asymmetry 2009 Volume 20(Issue 21) pp:2497-2503
Publication Date(Web):4 November 2009
DOI:10.1016/j.tetasy.2009.09.020
Co-reporter:Xiaoping Tang, Alexander J. Blake, William Lewis, Simon Woodward
Tetrahedron: Asymmetry 2009 Volume 20(Issue 16) pp:1881-1891
Publication Date(Web):26 August 2009
DOI:10.1016/j.tetasy.2009.07.006
Co-reporter:Antonella De Roma, Francesco Ruffo and Simon Woodward
Chemical Communications 2008 (Issue 42) pp:5384-5386
Publication Date(Web):17 Sep 2008
DOI:10.1039/B813137F
Modular phosphine ligands, synthesised rapidly from commercial N-acetylglucosamine, are very effective in copper(I)-catalysed 1,4-additions of ZnR2 to linear aliphatic enones (87–95% ee).
Co-reporter:Yvette Mata, Montserrat Diéguez, Oscar Pàmies, Simon Woodward
Inorganica Chimica Acta 2008 Volume 361(Issue 5) pp:1381-1384
Publication Date(Web):1 April 2008
DOI:10.1016/j.ica.2007.09.005
Modular sugar phosphite-oxazoline L1–L5a–c and phosphite-phosphoroamidite L6a–c ligand libraries were screened in the asymmetric Ni-catalyzed 1,2-addition reactions to aldehydes. Systematically varying the electronic and steric properties of the oxazoline and biaryl phosphite substituents and the functional groups attached to the basic sugar-backbone, we found a strong influence of the oxazoline and the functional groups of the sugar-backbone on the catalytic performance. Enantioselectivity (ee values up to 59%) was best with the catalysts precursor containing the phosphite-oxazoline ligand L3a, that contains a sterically hindered tert-butyl oxazoline group.Modular sugar phosphite-oxazoline and phosphite-phosphoroamidite ligand libraries were screened in the asymmetric Ni-catalyzed 1,2-addition reactions to aldehydes. Systematically varying the electronic and steric properties of the oxazoline and biaryl phosphite substituents and the functional groups attached to the basic sugar-backbone, we found a strong influence of the oxazoline and the functional groups of the sugar backbone on the catalytic performance.
Co-reporter:Kallolmay Biswas, Simon Woodward
Tetrahedron: Asymmetry 2008 Volume 19(Issue 14) pp:1702-1708
Publication Date(Web):25 July 2008
DOI:10.1016/j.tetasy.2008.06.017
Co-reporter:Jonathan Shannon, David Bernier, Daniel Rawson and Simon Woodward
Chemical Communications 2007 (Issue 38) pp:3945-3947
Publication Date(Web):15 Aug 2007
DOI:10.1039/B710681E
Addition of AlMe3 to commercial THF solutions of RZnX (R = aryl, functionalised aryl, vinyl; X = Br, I) simultaneously promotes Schlenk equilibria (leading to competent nucleophiles) and the formation of an Al–Zn-ligand catalyst delivering 80–90% ee for Ar1CH(OH)Ar2 formation from aldehydes.
Co-reporter:Claire Wilson;Rebecca E. Meadows;Maurizio Solinas;Alexer J. Blake and
European Journal of Organic Chemistry 2007 Volume 2007(Issue 10) pp:1613-1623
Publication Date(Web):9 FEB 2007
DOI:10.1002/ejoc.200600962
Reaction of 1,1′-binaphthol and structurally related 1,1′-biphenols with sulfonylating reagents [RSO2Cl; R = 4-Tol, Ph] leads to clean, ultraselective monoderivatisation, shown crystallographically in one case (2-OH–2′-OTs–1,1′-binaphthyl). Reaction of the remaining 2-hydroxy function with either Tf2O or NfF [Nf = nonaflate, CF3(CF2)3SO2-] affords the protected/activated cores (2-R1O–2′-R2O–1,1′-biaryl) [R1,R2 pairs = Tf,Ts (X-ray); Tf,SO2Ph; Nf,Ts (on 1,1′-binaphthyl core); Tf,Ts; Nf,Ts (on 1,1′-biphenyl core); Tf,Ts; Nf,Ts (on 3,3′,5,5′-tetramethyl-1,1′-biphenyl core)]. Reaction of the Tf,Ts species with either MeMgBr/NiCl2(dppe) (for the 1,1′-binaphthyl) or (AlMe3)2(DABCO)/Pd2(dba)3 (for the 1,1′-biphenyl) affords 2-OH–2′-Me–1,1′-biaryl units on subsequent hydrolysis (crystallographically characterised in the binaphthyl case). The latter methyl/hydroxy compounds are doubly deprotonated by either nBuLi/TMEDA or nBuLi/tBuOK to afford dianions that react cleanly with Cl2SiPh2 (two X-ray structures). The equivalent reaction of the 2-OH–2′-Me–1,1′-binaphthyl with Ph(O)Cl2 is less clean due to the absence of a strong Thorpe–Ingold effect. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2007)
Co-reporter:Alexer J. Blake ;Jonathan Shannon;John C. Stephens Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 9) pp:
Publication Date(Web):7 FEB 2007
DOI:10.1002/chem.200601739
The presence of promoted Schlenk equilibria for organozinc halide species has been explicitly demonstrated by 13C NMR studies. Thus, addition of methylaluminoxane (MeAlO)n, MAO, to RZnX (R=Et, Bn, ArCH2, (CH2)3CO2Et; X=Cl, Br) leads to the formation of ZnR2 and ZnX2⋅MAO. For EtZnCl, equilibration of ZnEt2 and ZnX2⋅MAO is rapid at −35 °C; a K value of 0.19 M−1 indicates the equilibrium favours ZnEt2 (0.75–3.0 equiv MAO). Use of RZnX/MAO mixtures allows copper-catalysed 1,4-addition to 2-cyclohexenone to be achieved, but a competing cascade reaction (two subsequent Michael additions and an intramolecular aldol reaction) leads to novel tetracyclic by-products (characterised crystallographically in one case). Activation of EtZnCl is also achieved by ZnMe2 addition and the presence of intermediate EtZnMe was observed by 13C NMR spectroscopy (at equilibrium, K≈1). Asymmetric conjugate addition in this system can be realised (up to 92 % ee for additions to 2-cyclohexenone).
Co-reporter:Thea Cooper;Andrew Novak;Luke D. Humphreys;Matthew D. Walker
Advanced Synthesis & Catalysis 2006 Volume 348(Issue 6) pp:
Publication Date(Web):5 APR 2006
DOI:10.1002/adsc.200505405
An extremely technically simple cross-methylation of aryl and vinyl halides and pseudohalides using an air-stable adduct of trimethylaluminium with a Pd(0) catalyst supported by commercially available biarylphosphines gives excellent yields of methylated products (mainly >95%). Reactions can be run with either 0.5 mol % catalyst or without requiring the exclusion of atmospheric oxygen or the drying of solvents in some cases. A wide variety of functional groups is tolerated including CN, OH, CO2R, CHO and NO2.
Co-reporter:Victoria E. Albrow;Alexer J. Blake;Ross Fryatt;Claire Wilson
European Journal of Organic Chemistry 2006 Volume 2006(Issue 11) pp:
Publication Date(Web):22 MAR 2006
DOI:10.1002/ejoc.200600071
Enantiopure lithiated 1,2-ferrocenes, [CpFe{1,2-η5-C5H3(Ar)(Li)}] (Ar = Ph, 4-CF3C6H4, 1-C10H7), react readily with PhOP(OR)2 to yield [CpFe{1,2-η5-C5H3(Ar)P(OR)2}] (R = Ph, 1,1′-biphenyl-based, 1,1′-binaphthyl-based) efficiently. Traditional routes to these species, involving the use of chlorophosphites ClP(OR)2 were found to be ineffective. These “ferrophite” ligands have been characterised by X-ray crystallography (4 examples) and shown to be effective in both nickel-catalysed addition of AlMe3 to PhCHO (up to 77 % ee) and copper(I)-catalysed additions of organoaluminium reagents to enones (up to 92 % ee). (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2006)
Co-reporter:Richard I. Robinson;Ross Fryatt;Claire Wilson
European Journal of Organic Chemistry 2006 Volume 2006(Issue 19) pp:
Publication Date(Web):2 AUG 2006
DOI:10.1002/ejoc.200600508
Literature N-alkylsaccharins (saccharin-R2) have been shown in some cases to be O-alkylated regioisomers by crystallography (3 structures). The genuine former species react with (S)-H2NCHR1CH2OH at 101 °C in dioxane to provide 1,2-C6H4(CONHCHR1CH2OH)(SO2NHR2) [R1 = H, Me, iPr, Bn, (CH2)2SMe; R2 = Bn, iPr, CHPh2, CHMePh]. The (iPr,Bn) compound is crystallographically characterised. If both R1 and R2 are sterically congested then reaction of the amino alcohol with the saccharin surrogate 1,2-C6H4(CO2Me)(SO2NHR2) is required. The saccharin-derived alcohols are converted into the oxazolines 1,2-C6H4(R1-oxazoline)(SO2NHR2) (R1 = H, Bn, Me, iPr; R2 = Bn, CHPh2, nPr, iPr, tBu, CHMePh). The dibenzyl compound is crystallographically characterised. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2006)
Co-reporter:Victoria Albrow, Alexander J. Blake, Alexandre Chapron, Claire Wilson, Simon Woodward
Inorganica Chimica Acta 2006 Volume 359(Issue 6) pp:1731-1742
Publication Date(Web):10 April 2006
DOI:10.1016/j.ica.2005.05.040
Ferrocene reacts with hexafluoroacetone trihydrate in refluxing octane to afford >80% yields of [CpFe(η5-C5H4C(CF3)2OH)] (X-ray), carrying out the reactions at 180 °C gives an additional 5% yield of [Fe(η5-C5H4C(CF3)2OH)2] (X-ray).The mono alcohol is lithiated with ButOK/BunLi/TMEDA affording partial conversion to mixtures of [CpFe(1,2-η5-C5H3C(CF3)2OH)(X)] and [Fe(η5-C5H4X)(1,2-η5-C5H3C(CF3)2OH)(X)] (X = SMe, CPh2OH) upon reaction with Me2S2 or OCPh2.For X = CPh2OH both structures are crystallographically characterised.Enantiopure [CpFe(1,2-η5-C5H3C(CF3)2OH)(SMe)] can be prepared from (R)-[CpFe(η5-C5H4S(O)C6H4Me)] via [CpFe(1,2-η5-C5H3S(O)C6H4Me)(C(CF3)2OH)] (X-ray) or [CpFe(1,2-η5-C5H3S(O)C6H4Me)(SMe)].Related procedures allow the preparation of [CpFe(1,2-η5-C5H3CPh2OH)(Y)] (Y = SMe, CHO (X-ray), C(CF3)2OH) and[CpFe(1,2-η5-C5H3C(CF3)2OH)(CHO)].The preparation of new ligand structures based on [CpFe(η5-C5H3(C(CF3)2OH)(E))] (E = SMe, CPh2OH, CHO). These are attained either by directed lithiation of [CpFe(η5-C5H4(C(CF3)2OH))] which results in the formation of the desired species and a [Fe(η5-C5H3(C(CF3)2OH(E)))(η5-C5H4E)] (E = SMe, CPh2OH) co-product or by sequential lithiation/trap reactions of [CpFe(η5-C5H4(S(O)4-Tol))].Crystal structures of examples of these complexes are presented.
Co-reporter:Alexandre Alexakis, Victoria Albrow, Kallolmay Biswas, Magali d'Augustin, Oscar Prieto and Simon Woodward
Chemical Communications 2005 (Issue 22) pp:2843-2845
Publication Date(Web):21 Apr 2005
DOI:10.1039/B503074A
Simple phosphoramidite ligands afford good to excellent levels of enantioselectivity in 1,4-additions of AlR3 species to enones; sequential carboalumination–ACA cascades are possible.
Co-reporter:Simon Woodward Dr.
Angewandte Chemie 2005 Volume 117(Issue 35) pp:
Publication Date(Web):29 JUL 2005
DOI:10.1002/ange.200501204
Das Handwerkszeug der Organokupferchemie ist nahezu komplett, seit kürzlich katalytische Methoden für ligandenvermittelte enantioselektive 1,4-konjugierte Additionen und SN2′-Substitutionen an Allylhalogeniden mit Grignard-, RZnCl- oder Organoaluminiumreagentien entwickelt wurden, die mit geringen Mengen an Kupfer(I)-Salzen auskommen (siehe Schema).
Co-reporter:Kallolmay Biswas Dr.;Oscar Prieto Dr.;Paul J. Goldsmith Dr.
Angewandte Chemie 2005 Volume 117(Issue 15) pp:
Publication Date(Web):14 MAR 2005
DOI:10.1002/ange.200462569
An Luft abgewogen! Das DABAL-Reagens (Me3Al)2⋅(DABCO) (DABCO=1,4-Diazabicyclo[2.2.2]octan) lässt sich einfach unter normalen Laborbedingungen handhaben, und chirale sekundäre Alkohole sind mit DABAL oder AlR3-Reagentien (R=Me, Et) effizient aus prochiralen Aldehyden zugänglich (siehe Schema; TOF=Umsatzfrequenz). Dies belegt, dass das DABAL-Reagens eine effiziente und bequeme Alternative für das Schumann-Blum-Reagens ist.
Co-reporter:Paul J. Goldsmith;Simon J. Teat Dr. Dr.
Angewandte Chemie 2005 Volume 117(Issue 15) pp:
Publication Date(Web):10 MAR 2005
DOI:10.1002/ange.200463028
Das Zink-Schlenk-Gleichgewicht, das seit seiner ersten Beschreibung 1966 wenig genutzt wurde, wird durch die Zugabe von Methylaluminoxan (MAO) begünstigt, das die Ausbeute an ZnR2 ausgehend von schädlichen RZnCl-Begleitprodukten maximiert. So wurde eine SN2′-Addition als hoch enantioselektiver Zugang zu chiralen β,β-disubstituierten α-Methylenpropionaten möglich (siehe Schema).
Co-reporter:Simon Woodward Dr.
Angewandte Chemie International Edition 2005 Volume 44(Issue 35) pp:
Publication Date(Web):29 JUL 2005
DOI:10.1002/anie.200501204
The organocopper craftsman's toolset is nearly complete with recent advances in ligand-based catalytic organocopper chemistry, which make viable enantioselective 1,4-conjugate additions and SN2′ substitutions of allylic halides using Grignard, RZnCl, and organoaluminum reagents (see scheme) at low loadings of copper(I) salts.
Co-reporter:Kallolmay Biswas Dr.;Oscar Prieto Dr.;Paul J. Goldsmith Dr.
Angewandte Chemie International Edition 2005 Volume 44(Issue 15) pp:
Publication Date(Web):14 MAR 2005
DOI:10.1002/anie.200462569
Weigh it out in air! The DABAL reagent (Me3Al)2⋅(DABCO) (DABCO=1,4-diazabicyclo[2.2.2]octane) can be easily handled under normal laboratory conditions. Furthermore, chiral secondary alcohols can be efficiently prepared from prochiral aldehydes (see scheme; TOF=turnover frequency) by using either DABAL or AlR3 reagents (R=Me, Et). Thus, DABAL can be used as an efficient, convenient alternative to the Schumann–Blum reagent.
Co-reporter:Paul J. Goldsmith;Simon J. Teat Dr. Dr.
Angewandte Chemie International Edition 2005 Volume 44(Issue 15) pp:
Publication Date(Web):10 MAR 2005
DOI:10.1002/anie.200463028
The zinc Schlenk equilibrium, little used since it was first described in 1966, has been promoted through the addition of methylaluminoxane (MAO) to maximize the yield of ZnR2 from deleterious RZnCl by-products. This process allows an SN2′-addition approach to the preparation of chiral β,β-disubstituted α-methylenepropionates with high enantioselectivity (see scheme).
Co-reporter:Anthony Cunningham, Vijaya Mokal-Parekh, Claire Wilson and Simon Woodward
Organic & Biomolecular Chemistry 2004 vol. 2(Issue 5) pp:741-748
Publication Date(Web):05 Feb 2004
DOI:10.1039/B313384B
In the presence of enantiopure MTBH2
(monothiobinaphthol, 2-hydroxy-2′mercapto-1,1′-binaphthyl; 0.2 eq.) quantitative allylation of ArC(O)Me takes place with impure Sn(CH2CHCH2)4
(prepared from allyl chloride, air-oxidised magnesium and SnCl4) to yield tert-homoallylic alcohols in 85–92% ee. In the same process highly purified, or commercial, Sn(CH2CHCH2)4 yields material of only 35–50% ee. The origin of these effects is the presence of small amounts of the compounds, EtSn(CH2CHCH2)3, ClSn(CH2CHCH2)3 ClSnEt(CH2CHCH2)2 in the tetraallyltin sample and the presence of traces of water (which inhibits achiral background reactions). All the triallyl and diallyl species enhance the stereoselectivity in the catalytic allylation reaction, the chlorides more so than the ethyl compound. Hydrolysis of ClSnEt(CH2CHCH2)2 affords crystallographically characterised Sn4(μ3-O)(μ2-Cl)2Cl2Et4(CH2CHCH2)4. Reaction of this latter compound with MTBH2 leads to the most potent catalyst.
Co-reporter:Richard I. Robinson;John C. Stephens;Steve M. Worden;Alexer J. Blake;Claire Wilson
European Journal of Organic Chemistry 2004 Volume 2004(Issue 22) pp:
Publication Date(Web):2 NOV 2004
DOI:10.1002/ejoc.200400562
2-Sulfobenzoic acid anhydride opens cleanly with N-Boc-protected α-amino alcohols to afford zwitterionic esters 1,2-C6H4(SO3−)(CO2CH2CHR1NH3+) (R1 = H, alkyl), the species R1 = iPr is crystallographically characterised. Dehydration of these species affords zwitterionic oxazoline sulfonic acid derivatives 1,2-C6H4(SO3−)(CA=NH+CHR1CH2OA) (CA and OA are bonded C−O) for R1 = iPr (X-ray), iBu, tBu, CH2Ph. Reaction with water regenerates the zwitterionic esters while exposure to M(OAc)2 (M = Cu, Pd) leads to the formation of the crystallographically characterised complexes MIIL2 (L = anion of the iPr-oxazolinesulfonic acid). Heating either the zwitterionic esters or oxazolines with R2NH2 [R2 = CH2Ph, CH2(1-C10H7), R-CH(Me)Ph] leads to SN2 attack at the oxazoline methylene group leading to the amido sulfonic acids 1,2-C6H4(CONHCHRCH2NHR2)(SO3H) based on crystallographic studies on R1 = iPr, R2 = CH2(1-C10H7). (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2004)
Co-reporter:Paul K. Fraser Dr.
Chemistry - A European Journal 2003 Volume 9(Issue 3) pp:
Publication Date(Web):29 JAN 2003
DOI:10.1002/chem.200390087
2-Hydroxy-2′-alkylthio-1,1′-binaphthyl compounds are catalytic promoters of the 1,4-addition of AlMe3 to linear aliphatic enones in THF at −40 to −48 °C in the presence of [Cu(MeCN)4]BF4. At ligand loadings of 5–20 mol %, enantioselectivities of 80–93 % are realised for most substrates. To attain these values, the use of highly pure AlMe3 is mandatory. The presence of methylalumoxane (MAO), derived by hydrolysis, leads to reduced enantioselectivity and a conjugate addition product.
Co-reporter:Victoria Albrow, Kallolmay Biswas, Andrew Crane, Nicholas Chaplin, Tim Easun, Serafino Gladiali, Barry Lygo, Simon Woodward
Tetrahedron: Asymmetry 2003 Volume 14(Issue 18) pp:2813-2819
Publication Date(Web):19 September 2003
DOI:10.1016/S0957-4166(03)00583-4
Acylation of octahydro-BINOL with Me2NC(S)Cl, in the presence of NaH, allows the formation of 2-(OH)-2′-(Me2NC(S)O)-1,1′-C10H20. Subsequent Newman–Kwart rearrangement (275°C, 12 min) proceeds cleanly with a small amount of racemisation (96–97% ee). The equivalent BINOL-derived species undergoes an identical rearrangement (but with higher racemisation, 79–80%) and appreciable amounts of a thiophene by-product are formed. Semi-empirical calculations (PM3) predict a higher racemisation barrier for the octahydro compound and suggest that the Newman–Kwart rearrangement could proceed via a concerted pathway. The H8-BINOL derived compound can be modified to 2-(OH)-2′-(SBun)-1,1′-C10H20 and the BINOL species to 2-(OH)-2′-(SBut)-1,1′-C10H12. The former promotes the 1,4-addition of AlMe3 to non-3-en-2-one in 57% ee the latter in 63% ee.Graphic(−)-(Sa)-2-(N,N-Dimethylcarbamoylthio)-2′-hydroxy-1,1′-binaphthylC23H19NO2SE.e.=80%[α]20D=−277 (c 0.52, CHCl3)Source of chirality: (Sa)-BINOLAbsolute configuration: Sa(−)-(Sa)-2-S-t-Butyl-2′-hydroxymercapto-1,1′-binaphthylC24H22OSE.e.=80%[α]20D=−32 (c 0.47, CHCl3)Source of chirality: (Sa)-BINOLAbsolute configuration: Sa(−)-(Sa)-2-(N,N-Dimethylthiocarbamoyloxy)-2′-hydroxy-5,5′,6,6′,7,7′,8,8′-octahydro-1,1′-binaphthylC23H27NO2SE.e. >98%[α]20D=−208 (c 0.53, CHCl3)Source of chirality: (Sa)-BINOLAbsolute configuration: Sa(−)-(Sa)-2-(N,N-Dimethylcarbamoylthio)-2′-hydroxy-5,5′,6,6′,7,7′,8,8′-octahydro-1,1′-binaphthylC23H27NO2SE.e.=97%[α]29D=−155 (c 0.93, CHCl3)Source of chirality: (Sa)-BINOLAbsolute configuration: Sa(−)-(Sa)-2-Hydroxy-2′-mercapto-5,5′,6,6′,7,7′,8,8′-octahydro-1,1′-binaphthylC20H22OSE.e.=97%[α]20D=−73 (c 0.56, CHCl3)Source of chirality: (Sa)-BINOLAbsolute configuration: Sa(−)-(Sa)-2-S-n-Butyl-2′-hydroxymercapto-5,5′,6,6′,7,7′,8,8′-octahydro-1,1′-binaphthylC24H30OSE.e.=97%[α]20D=−30 (c 0.57, CHCl3)Source of chirality: (Sa)-BINOLAbsolute configuration: Sa
Co-reporter:Elizabeth J. MacLean, Richard I. Robinson, Simon J. Teat, Claire Wilson and Simon Woodward
Dalton Transactions 2002 (Issue 18) pp:3518-3524
Publication Date(Web):14 Aug 2002
DOI:10.1039/B205548A
Reaction of 4,4′,4″-tri(tert-butyl)-2,2′:6′,2″-terpyridine (terpy*) with gaseous SO2–O2 mixtures and [PdCl2(MeCN)2] followed by aqueous workup affords crystallographically characterised [PdCl(terpy*)]Cl and both O- and S-bound [Pd(SO3)(terpy*)]. The former is also available from [PdCl2(MeCN)n]
(n
= 0, 2) and terpy*; the latter from aerobic oxidation of [Pd(dba)2]–terpy*–SO2 mixtures (dba = dibenzylideneacetone). Oxidation of [Pd(SO3)(terpy*)] with O2
(at 210 °C in PhNO2) yields crystallographically characterised O-bound [Pd(SO4)(terpy*)]. The crystal structures of two related compounds are also reported: [Pd(OAc)(terpy*)]Cl and [PdCl2(bipy*)].
Co-reporter:Thomas von Schrader
European Journal of Organic Chemistry 2002 Volume 2002(Issue 22) pp:
Publication Date(Web):28 OCT 2002
DOI:10.1002/1099-0690(200211)2002:22<3833::AID-EJOC3833>3.0.CO;2-K
Naphthalene-catalysed reductions of PhCH(OR)2 (R = Me, CH2CH2OMe) acetals by lithium metal, followed by reactions with electrophiles (H+, TMSCl, nBuBr, CH2=CHCH2Br), proceed with high chemoselectivity when the reductions are carried out at −90 °C especially for R = CH2CH2OMe. (© Wiley-VCH Verlag GmbH, 69451 Weinheim, Germany, 2002)
Co-reporter:Victor Garcia-Ruiz, Simon Woodward
Tetrahedron: Asymmetry 2002 Volume 13(Issue 19) pp:2177-2180
Publication Date(Web):4 October 2002
DOI:10.1016/S0957-4166(02)00574-8
The alcohols (S)-C5H11nCH(R)CH2OH (R=Me, Et) have been prepared by Evans’ alkylation chemistry (>98% e.e.). For R=Me [α]D=−15.5 (c 0.31, MeOH); for R=Et [α]D=+6.8 (c 0.31, MeOH). Equivalent alcohols are obtained by Baeyer–Villiger oxidative cleavage of (S)-(−)-C5H11nCH(R)CH2COMe (R=Me, 85% e.e.; R=Et, 62% e.e.) derived from catalytic asymmetric conjugate addition. Thus, AlMe3 or ZnEt2 addition to the Si face of the enone generates a (−) antipode with a 4S stereocentre.Graphic(S)-4-Methylnonan-2-oneC10H20OE.e.=85%[α]D=−3.3 (c 0.31, MeOH)Source of chirality: catalytic asymmetric conjugate additionAbsolute configuration: 4S(S)-4-Ethylnonan-2-oneC11H22OE.e.=62%[α]D=−2.4 (c 0.31, MeOH)Source of chirality: catalytic asymmetric conjugate additionAbsolute configuration: 4S(S)-2-Methylheptan-1-olC8H18OE.e. >98%[α]D=−15.5 (c 0.31, MeOH)Source of chirality: alkylation of Evans’ auxiliary followed by LiAlH4Absolute configuration: 2S(S)-2-Ethylheptan-1-olC9H20OE.e. >98%[α]D=+6.8 (c 0.31, MeOH)Source of chirality: alkylation of Evans’ auxiliary followed by LiAlH4Absolute configuration: 2S(4S)-Benzyl-3-((2S)-methylheptanoyl)oxazolidin-2-oneC18H25NO3E.e. >98%[α]D=+140.0 (c 0.30, Et2O)Source of chirality: alkalation of Evans’ enolateAbsolute configuration: 2S,4S(4S)-Benzyl-3-((2S)-ethylheptanoyl)oxazolidin-2-oneC19H27NO3E.e. >98%[α]D=+126.7 (c 0.30, Et2O)Source of chirality: alkylation of Evans’ enolateAbsolute configuration: 2S,4S
Co-reporter:Christoph Börner;Michael R. Dennis;Ekkehard Sinn
European Journal of Organic Chemistry 2001 Volume 2001(Issue 13) pp:
Publication Date(Web):5 JUN 2001
DOI:10.1002/1099-0690(200107)2001:13<2435::AID-EJOC2435>3.0.CO;2-0
Directed ortho dilithiation of bis(diethylcarbamate) or bis(MOM)-protected (Sa)-1,1′-bi(2-naphthol) followed by treatment with R2S2 [R = Me, Ph (X-ray structure)] or Me2Se2 cleanly affords the 3,3′ derivatives; the free naphthols are produced on deprotection. In the case of the bis(MOM) series, but not that of the bis(carbamates), some racemisation occurs. The ligand 2,2′-dihydroxy-3,3′-dimethylthio-1,1′-binaphthalene shows optimal performance in the addition of ZnEt2 to linear aliphatic enones (E)-R1C(O)CH=CHR2. Variation of the steric demands of R1 and R2 generates catalytic results consistent with binding of a zinc-based Lewis acid anti to the ene function and with the reactive conformation being s-cis. With enones containing the functions R2 = (CH2)nCH(OAlkyl)2 (n = 0−2), the ZnEt2 addition products undergo base-promoted cyclisation.
Co-reporter:Christoph Börner, Josep Gimeno, Serafino Gladiali, Paul J. Goldsmith, Daniela Ramazzotti and Simon Woodward
Chemical Communications 2000 (Issue 24) pp:2433-2434
Publication Date(Web):22 Nov 2000
DOI:10.1039/B006943O
The copper-catalysed SN2′ addition of
ZnR2 to allylic
(Z)-ArCHC(CH2X)(CO2Et) (X = Br, Cl,
OSO2Me) fashions only
ArCH(R)C(CH2)(CO2Et); use of a chiral ligand
gives up to 64% ee for this demanding reaction.
Co-reporter:Alexer J. Blake Dr.;Anthony Cunningham;Alan Ford Dr.;Simon J. Teat Dr. Dr.
Chemistry - A European Journal 2000 Volume 6(Issue 19) pp:
Publication Date(Web):18 SEP 2000
DOI:10.1002/1521-3765(20001002)6:19<3586::AID-CHEM3586>3.0.CO;2-S
LiGaH4, in combination with the S,O-chelate 2-hydroxy-2′-mercapto-1,1′-binaphthyl (MTBH2), forms an active catalyst for the asymmetric reduction of prochiral ketones, with catecholborane as the hydride source. Enantioface differentiation is on the basis of the steric requirements of the ketone substituents. Aryl/n-alkyl ketones are reduced in 90–93 % ee and RC(O)Me (e.g. R=iPr, cycloC6H11, tBu) in 60–72 % ee. Other borane sources and alternative catalyst structures based on indium do not form enantioselective catalysts.
Co-reporter:Jonathan Shannon, David Bernier, Daniel Rawson and Simon Woodward
Chemical Communications 2007(Issue 38) pp:NaN3947-3947
Publication Date(Web):2007/08/15
DOI:10.1039/B710681E
Addition of AlMe3 to commercial THF solutions of RZnX (R = aryl, functionalised aryl, vinyl; X = Br, I) simultaneously promotes Schlenk equilibria (leading to competent nucleophiles) and the formation of an Al–Zn-ligand catalyst delivering 80–90% ee for Ar1CH(OH)Ar2 formation from aldehydes.
Co-reporter:Antonella De Roma, Francesco Ruffo and Simon Woodward
Chemical Communications 2008(Issue 42) pp:NaN5386-5386
Publication Date(Web):2008/09/17
DOI:10.1039/B813137F
Modular phosphine ligands, synthesised rapidly from commercial N-acetylglucosamine, are very effective in copper(I)-catalysed 1,4-additions of ZnR2 to linear aliphatic enones (87–95% ee).
Co-reporter:D. Willcox, S. Woodward and A. Alexakis
Chemical Communications 2014 - vol. 50(Issue 14) pp:NaN1657-1657
Publication Date(Web):2014/01/07
DOI:10.1039/C3CC48191C
Chloromethylvinyl alanes (E)-ClMeAl(CHCHR) prepared directly from terminal alkynes undergo 1,4-addition to cyclohexenone and 3-methylcyclohexenone in moderate to good yield (30–70%) and good to excellent stereoselectivity (80–98% ee) using readily available copper(I) sources and chiral ligands.
Co-reporter:Philip Andrews, Christopher M. Latham, Marc Magre, Darren Willcox and Simon Woodward
Chemical Communications 2013 - vol. 49(Issue 15) pp:NaN1490-1490
Publication Date(Web):2013/01/14
DOI:10.1039/C2CC37537K
Stabilized AlHCl2·(THF)2 hydroaluminates RCCH with exceptional chemo-, regio- and stereoselectivity under efficient ZrCl2(η-C5Me5)2 catalysis (2–5 mol%). The resulting vinyl alanes undergo palladium cross-coupling with a wide range of sp2 electrophiles (aryl, heteroaryl and vinyl halides/pseudohalides) in good to excellent yields.