Co-reporter:Yifei Li, Kelvin E. Jackson, Andrew Charlton, Ben Le Neve-Foster, Asma Khurshid, Heinrich-K. A. Rudy, Amber L. Thompson, Robert S. Paton, and David. M. Hodgson
The Journal of Organic Chemistry October 6, 2017 Volume 82(Issue 19) pp:10479-10479
Publication Date(Web):September 27, 2017
DOI:10.1021/acs.joc.7b01954
Quantum chemical studies of C-ethylation of 1-methyl- and 1,4,4-trimethyl-tropane-derived enamines predict good (89:11 er, B3LYP) and high (98:2 er, B3LYP) levels, respectively, of asymmetric induction in the resulting α-alkylated aldehydes. The nonracemic tropanes were synthesized using Mannich cyclization strategies (Robinson-Schöpf and by way of a Davis-type N-sulfinyl amino bisketal, respectively), and ethylation of the derived enamines was found to support the predicted sense and magnitude of asymmetric induction (81:19 er and 95:5 er, respectively). A comparison of several computational methods highlights the robustness of predicted trends in enantioselectivity, enabling theory to guide synthesis.
Co-reporter:Aubert Ribaucourt and David M. Hodgson
Organic Letters 2016 Volume 18(Issue 17) pp:4364-4367
Publication Date(Web):August 16, 2016
DOI:10.1021/acs.orglett.6b02120
The sodium salts E-15 and Z-15 of the originally proposed dihydropyran acid structure of aruncin B (1) were prepared through ring-closing alkene metathesis (RCM) and ethoxyselenation–selenoxide elimination, but acid sensitivity of these salts, together with inconsistencies in the spectral data, suggested a significant structural misassignment. A β-iodo Morita–Baylis–Hillman reaction to give Z-iodo ester 24, followed by Sonogashira cross-coupling–5-exo-dig lactonization, provided concise access to a Z-γ-alkylidenebutenolide 18, which possessed data corresponding to those originally reported for aruncin B.
Co-reporter:Reece Jacques, Ritashree Pal, Nicholas A. Parker, Claire E. Sear, Peter W. Smith, Aubert Ribaucourt and David M. Hodgson
Organic & Biomolecular Chemistry 2016 vol. 14(Issue 25) pp:5875-5893
Publication Date(Web):25 Apr 2016
DOI:10.1039/C6OB00593D
In the past two decades, alkene metathesis has risen in prominence to become a significant synthetic strategy for alkene formation. Many total syntheses of natural products have used this transformation. We review the use, from 2003 to 2015, of ring-closing alkene metathesis (RCM) for the generation of dihydrofurans or -pyrans in natural product synthesis. The strategies used to assemble the RCM precursors and the subsequent use of the newly formed unsaturation will also be highlighted and placed in context.
Co-reporter:David M. Hodgson, Claire L. Mortimer, and Jeffrey M. McKenna
Organic Letters 2015 Volume 17(Issue 2) pp:330-333
Publication Date(Web):December 23, 2014
DOI:10.1021/ol503441d
tert-Butoxythiocarbonyl (Botc), the long-neglected thiocarbonyl analogue of Boc, facilitates (unlike its alkoxycarbonyl cousin) α-lithiation and electrophile incorporation on N-Botc-azetidine. N,N,N′,N′-endo,endo-Tetramethyl-2,5-diaminonorbornane proved optimal as a chiral ligand, generating adducts with er up to 92:8. Facile deprotection, under conditions that left the corresponding N-Boc systems intact, was achieved using either TFA or via thermolysis in ethanol.
Co-reporter:Kelvin E. Jackson, Claire L. Mortimer, Barbara Odell, Jeffrey M. McKenna, Timothy D. W. Claridge, Robert S. Paton, and David M. Hodgson
The Journal of Organic Chemistry 2015 Volume 80(Issue 20) pp:9838-9846
Publication Date(Web):September 24, 2015
DOI:10.1021/acs.joc.5b01804
1H NMR and computational analyses provide insight into the regiodivergent (α- and α′-) lithiation–electrophile trapping of N-thiopivaloyl- and N-(tert-butoxythiocarbonyl)-α-alkylazetidines. The magnitudes of the rotation barriers in these azetidines indicate that rotamer interconversions do not occur at the temperature and on the time scale of the lithiations. The NMR and computational studies support the origin of regioselectivity as being thiocarbonyl-directed lithiation from the lowest energy amide-like rotameric forms (cis for N-thiopivaloyl and trans for N-tert-butoxythiocarbonyl).
Co-reporter:David M. Hodgson, Christopher I. Pearson, and Madiha Kazmi
Organic Letters 2014 Volume 16(Issue 3) pp:856-859
Publication Date(Web):January 10, 2014
DOI:10.1021/ol403626k
s-BuLi-induced α-lithiation–elimination of LiOMe from N-Boc-3-methoxyazetidine and further in situ α-lithiation generates N-Boc-2-lithio-2-azetine which can be trapped with electrophiles, either directly (carbonyl or heteroatom electrophiles) or after transmetalation to copper (allowing allylations and propargylations), providing a concise access to 2-substituted 2-azetines.
Co-reporter:David M. Hodgson, Andrew Charlton
Tetrahedron 2014 70(13) pp: 2207-2236
Publication Date(Web):
DOI:10.1016/j.tet.2013.11.046
Co-reporter:David M. Hodgson, Stanislav Man, Kimberley J. Powell, Ziga Perko, Minxiang Zeng, Elena Moreno-Clavijo, Amber L. Thompson, and Michael D. Moore
The Journal of Organic Chemistry 2014 Volume 79(Issue 20) pp:9728-9734
Publication Date(Web):October 3, 2014
DOI:10.1021/jo501893r
Rh(II)-catalyzed oxonium ylide formation–[2,3] sigmatropic rearrangement of α-diazo-β-ketoesters possessing γ-cyclic unsaturated acetal substitution, followed by acid-catalyzed elimination–lactonization, provides a concise approach to 1,7-dioxaspiro[4.4]non-2-ene-4,6-diones. The process creates adjacent quaternary stereocenters with full control of the relative stereochemistry. An unsymmetrical monomethylated cyclic unsaturated acetal leads to hyperolactone C, where ylide formation–rearrangement proceeds with high selectivity between subtly nonequivalent acetal oxygen atoms.
Co-reporter:David M. Hodgson, Elena Moreno-Clavijo, Sophie E. Day and Stanislav Man
Organic & Biomolecular Chemistry 2013 vol. 11(Issue 32) pp:5362-5369
Publication Date(Web):12 Jul 2013
DOI:10.1039/C3OB41251B
A stereocontrolled synthesis of norphenyl hyperolactone C together with its utility as a direct precursor to the anti-HIV agent hyperolactone C and analogues by addition of organolithiums, are described. Preliminary studies to access this key building block in a catalytic enantioselective manner are also reported.
Co-reporter:David M. Hodgson, Andrew Charlton, Robert S. Paton, and Amber L. Thompson
The Journal of Organic Chemistry 2013 Volume 78(Issue 4) pp:1508-1518
Publication Date(Web):January 29, 2013
DOI:10.1021/jo3025972
The synthesis and alkylation of chiral, nonracemic tropane- and homotropane-derived enamines is examined as an approach to enantioenriched α-alkylated aldehydes. The two bicyclic N auxiliaries, which differ by a single methylene group, give opposite senses of asymmetric induction on alkylation with EtI and provide modestly enantioenriched 2-ethylhexanal (following hydrolysis of the alkylated iminium). The observed stereoselectivity is supported by density functional studies of ethylation for both enamines.
Co-reporter:David M. Hodgson, Christopher I. Pearson, and Amber L. Thompson
The Journal of Organic Chemistry 2013 Volume 78(Issue 3) pp:1098-1106
Publication Date(Web):January 10, 2013
DOI:10.1021/jo3025225
α-Lithiation of N-thiopivaloylazetidin-3-ol and subsequent electrophile trapping provides access to a range of 2-substituted 3-hydroxyazetidines with generally good trans-diastereoselectivity, aside from deuteration, which gives the cis-diastereoisomer. Deuterium labeling studies indicate that the initial α-deprotonation occurs preferentially, but not exclusively, in a trans-selective manner. These studies also suggest that the stereochemical outcome of the electrophile trapping depends on the electrophile used but is independent of which α-proton (cis or trans to the hydroxyl group) is initially removed.
Co-reporter:David M. Hodgson and Saifullah Salik
Organic Letters 2012 Volume 14(Issue 17) pp:4402-4405
Publication Date(Web):August 16, 2012
DOI:10.1021/ol3018853
Lithiation–in situ silylation of terminal epoxides using lithium 2,2,6,6-tetramethylpiperidide in combination with phenyldimethyl(or diethyl)silyl chloride provides a direct process for the synthesis of trans-α,β-epoxysilanes, which undergo α-ring opening with alkenylcoppers to give syn-β-hydroxyallylic silanes. The chemistry is applied in an annulation approach to the C10–C19 tetrahydrofuran-containing portion of the lytophilippines.
Co-reporter:David M. Hodgson, Eric P. A. Talbot and Barry P. Clark
Chemical Communications 2012 vol. 48(Issue 51) pp:6349-6350
Publication Date(Web):22 May 2012
DOI:10.1039/C2CC32283H
The first synthesis of the bioactive sesquiterpene lactone hydroxyanthecotulide is achieved in 7 steps, involving a stereocontrolled Cr(II)-catalysed reaction of 3-(bromomethyl)furan-2(5H)-one with enynal 9 and a mild Au(I)-catalysed Meyer–Schuster rearrangement.
Co-reporter:David M. Hodgson and Rosanne S. D. Persaud
Organic & Biomolecular Chemistry 2012 vol. 10(Issue 39) pp:7949-7951
Publication Date(Web):30 Aug 2012
DOI:10.1039/C2OB26346G
β-Lithiooxyphosphonium ylides, made in situ from an aldehyde and methylenetriphenylphosphorane, react with a second aldehyde to form E-allylic alcohols. α-Branching and α,β-unsaturation in the second aldehyde, together with the lack of further substitution on the phosphorane carbon play important roles in selectivity. A range of these aldehydes, in addition to aromatic aldehydes as the second aldehyde also provided synthetically useful access to E-allylic alcohols.
Co-reporter:David M. Hodgson and Tanzeel Arif
Chemical Communications 2011 vol. 47(Issue 9) pp:2685-2687
Publication Date(Web):12 Jan 2011
DOI:10.1039/C0CC04429F
β-Lithiooxyphosphonium ylides, generated in situ from aldehydes and methylenetriphenylphosphorane, react with halomethyl esters to form disubstituted allylic esters in good yields and with high Z-selectivity.
Co-reporter:David M. Hodgson, Eric P. A. Talbot, and Barry P. Clark
Organic Letters 2011 Volume 13(Issue 10) pp:2594-2597
Publication Date(Web):April 13, 2011
DOI:10.1021/ol200711f
Zinc or a chromium(II) source with 3-(bromomethyl)furan-2(5H)-one (3) and an aldehyde gives β-(hydroxymethylaryl/alkyl)-α-methylene-γ-butyrolactones 5 in good yields and high diastereoselectivities. The methodology is demonstrated in concise syntheses of (±)-hydroxymatairesinol (8) and (±)-methylenolactocin (10) by subsequent arylboronate conjugate addition and translactonization, respectively.
Co-reporter: David M. Hodgson;Dr. Stanislav Man
Chemistry - A European Journal 2011 Volume 17( Issue 35) pp:9731-9737
Publication Date(Web):
DOI:10.1002/chem.201101082
Abstract
Starting from readily available (S)-styrene oxide an asymmetric synthesis is described of the naturally occurring anti-HIV spirolactone (−)-hyperolactone C, which possesses adjacent fully substituted stereocenters. The key step involves a stereocontrolled RhII-catalysed oxonium ylide formation–[2,3] sigmatropic rearrangement of an α-diazo-β-ketoester bearing allylic ether functionality. From the resulting furanone, an acid-catalysed lactonisation and dehydrogenation gives the natural product.
Co-reporter:David M. Hodgson and Tanzeel Arif
Organic Letters 2010 Volume 12(Issue 18) pp:4204-4207
Publication Date(Web):August 19, 2010
DOI:10.1021/ol101843q
β-Lithiooxyphosphonium ylides, generated in situ from aldehydes and Wittig reagents, react readily with halomethyl esters to form trisubstituted Z-allylic esters. The methodology was applied to a total synthesis of the geranylgeraniol-derived diterpene (6S,7R,Z)-7-hydroxy-2-((E)-6-hydroxy-4-methylhex-4-enylidene)-6,10-dimethylundec-9-enyl acetate (12).
Co-reporter:David M. Hodgson, Ruth E. Shelton, Thomas A. Moss and Mouloud Dekhane
Organic Letters 2010 Volume 12(Issue 12) pp:2834-2837
Publication Date(Web):May 14, 2010
DOI:10.1021/ol100943j
An efficient Lewis acid induced nitrogen-driven rearrangement iminium-trapping cascade from an epoxytropinone 3 gives a 7-allylated 6-azabicyclo[3.2.1]octan-3-one 2, which is converted into the alkaloid (±)-peduncularine (1).
Co-reporter:David M. Hodgson, Carolina Villalonga-Barber, Jonathan M. Goodman and Silvina C. Pellegrinet
Organic & Biomolecular Chemistry 2010 vol. 8(Issue 17) pp:3975-3984
Publication Date(Web):02 Jul 2010
DOI:10.1039/C004496B
Reaction of diazodiketoesters 17 and 28 with methyl glyoxylate in the presence of catalytic rhodium(II) acetate generates predominantly the 6,8-dioxabicyclo[3.2.1]octanes 29 and 30, respectively. Acid-catalysed rearrangement of the corresponding alcohol 31 favours, at equilibrium, the 2,8-dioxabicyclo[3.2.1]octane skeleton 33 of the squalestatins–zaragozic acids. Force field calculations on the position of the equilibrium gave misleading results. DFT calculations were correct in suggesting that the energy difference between 31 and 33 should be small, but did not always suggest the right major product. Calculation of the NMR spectra of the similar structures could be used to assign the isomers with a high level of confidence.
Co-reporter:JoannM. Um;NaeemS. Kaka Dr.;DavidM. Hodgson ;K.N. Houk
Chemistry - A European Journal 2010 Volume 16( Issue 21) pp:6310-6316
Publication Date(Web):
DOI:10.1002/chem.201000046
Abstract
The asymmetric C-alkylation of chiral enamines derived from terminal epoxides and lithium 2,2,6-trialkylpiperidides has previously been shown to provide α-alkylated aldehydes by intermolecular nucleophilic substitution in good levels of asymmetric induction. We now report a computational study of the origins of asymmetric induction in these reactions. Computational modeling with density functional theory (B3LYP/6-31G(d)) agrees closely with the experimental observations. This stereoselectivity is attributed to a preferential conformation of the enamine and the piperidine ring that places the C-6 alkyl substituent in an axial position due to A1, 3 strain. Preferential attack occurs away from the axial group, for steric reasons. The effects of changing the C-6 substituent from methyl to isopropyl were studied, and twist transition states were found to contribute significantly in the latter alkylations.
Co-reporter:DavidM. Hodgson ;Johannes Kloesges
Angewandte Chemie International Edition 2010 Volume 49( Issue 16) pp:2900-2903
Publication Date(Web):
DOI:10.1002/anie.201000058
Co-reporter:DavidM. Hodgson ;Johannes Kloesges
Angewandte Chemie 2010 Volume 122( Issue 16) pp:2962-2965
Publication Date(Web):
DOI:10.1002/ange.201000058
Co-reporter:David M. Hodgson, Matthew L. Jones, Christopher R. Maxwell, Andrew R. Cowley, Amber L. Thompson, Osamu Ichihara, Ian R. Matthews
Tetrahedron 2009 65(37) pp: 7825-7836
Publication Date(Web):
DOI:10.1016/j.tet.2009.07.022
Co-reporter:David M. Hodgson, Rebecca Glen, Alison J. Redgrave
Tetrahedron: Asymmetry 2009 Volume 20(6–8) pp:754-757
Publication Date(Web):7 May 2009
DOI:10.1016/j.tetasy.2009.02.031
The synthesis and 1,3-dipolar cycloaddition reactions of unsaturated α-diazo-β,ε-diketo sulfones 7 using chiral rhodium catalysts (up to 71.5:28.5 er) are described.7-(Phenylsulfonyl)-11-oxatricyclo[5.3.1.01,5]undecan-8-oneC16H18O4Ser = 71.5:28.5 (by chiral HPLC)[α]D25=+13.6 (c 1.0, CH2C12)Source of chirality: Rh (S)-valine-derived catalystAbsolute configuration: unknown7-(Methylsulfonyl)-11-oxatricyclo[5.3.1.01,5]undecan-8-oneC11H16O4Ser = 66.5:33.5 (by chiral GC)[α]D25=+3.1 (c 1.05, CH2C12)Source of chirality: Rh (S)-valine-derived catalystAbsolute configuration: unknown
Co-reporter:DavidM. Hodgson ;NaeemS. Kaka
Angewandte Chemie 2008 Volume 120( Issue 51) pp:10106-10108
Publication Date(Web):
DOI:10.1002/ange.200804369
Co-reporter:DavidM. Hodgson ;NaeemS. Kaka
Angewandte Chemie International Edition 2008 Volume 47( Issue 51) pp:9958-9960
Publication Date(Web):
DOI:10.1002/anie.200804369
Co-reporter:David M. Hodgson and Leonard H. Winning
Organic & Biomolecular Chemistry 2007 vol. 5(Issue 19) pp:3071-3082
Publication Date(Web):02 Aug 2007
DOI:10.1039/B707566A
Radical rearrangements are important transformations in organic synthesis. The stabilisation of α-nitrogen radicals is shown to be a useful effect for the control of radical rearrangements and is applied to the synthesis of a variety of azabicyclic frameworks. The utility of this method is illustrated in the synthesis of bioactive targets.
Co-reporter:David M. Hodgson ;Philip G. Humphreys;Zhaoqing Xu Dr.;John G. Ward Dr.
Angewandte Chemie 2007 Volume 119(Issue 13) pp:
Publication Date(Web):15 FEB 2007
DOI:10.1002/ange.200604920
Einfach entschützt: Lithium-2,2,6,6-tetramethylpiperidid löst in Aziridinen mit N-Boc- oder N-Phosphonat-Schutzgruppen eine regio- und stereoselektive Wanderung der Schutzgruppe vom N- zum C-Atom aus. Die Produkte sind präparativ nützliche trans-Aziridinylester bzw. trans-Aziridinylphosphonate.
Co-reporter:David M. Hodgson ;Deepshikha Angrish
Chemistry - A European Journal 2007 Volume 13(Issue 12) pp:
Publication Date(Web):9 FEB 2007
DOI:10.1002/chem.200601692
Highly stereoselective formation of cis-2-ene-1,4-diesters by homo- and heterocoupling of α-diazoacetates in the presence of Grubbs second-generation catalyst is demonstrated. The dual reactivity of the catalyst in alkene metathesis and diazocoupling has been exploited in the synthesis of 12–26-membered macrocyclic dienyl dilactones by one-pot carbene dimerisation/ring-closing metathesis.
Co-reporter:David M. Hodgson ;Philip G. Humphreys;Zhaoqing Xu Dr.;John G. Ward Dr.
Angewandte Chemie International Edition 2007 Volume 46(Issue 13) pp:
Publication Date(Web):15 FEB 2007
DOI:10.1002/anie.200604920
Benefiting from deprotection: Lithium 2,2,6,6-tetramethylpiperidide induces N-Boc or N-phosphonate terminal aziridines to undergo regio- and stereoselective N-to-C migration of the protecting group, giving synthetically valuable trans-aziridinylesters and trans-aziridinylphosphonates.
Co-reporter:David M. Hodgson, Deepshikha Angrish and Agnès H. Labande
Chemical Communications 2006 (Issue 6) pp:627-628
Publication Date(Web):18 Jan 2006
DOI:10.1039/B515943A
Dicarbonyl-stabilised diazo functionality is tolerated during alkene cross-metathesis using Grubbs' catalyst, but undergoes subsequent tandem carbonyl ylide formation–intramolecular 1,3-dipolar cycloaddition on addition of catalytic Rh2(OAc)4 in a one-pot operation.
Co-reporter:David M. Hodgson, Matthew J. Fleming, Zhaoqing Xu, Changxue Lin and Steven J. Stanway
Chemical Communications 2006 (Issue 30) pp:3226-3228
Publication Date(Web):22 Jun 2006
DOI:10.1039/B606583J
N-Tosyl-protected 3-hydroxypyrrolidines are prepared by reaction of dimethylsulfoxonium methylide with readily available epoxysulfonamides.
Co-reporter:David M. Hodgson;Deepshikha Angrish
Advanced Synthesis & Catalysis 2006 Volume 348(Issue 16-17) pp:
Publication Date(Web):27 NOV 2006
DOI:10.1002/adsc.200600306
Chemoselective cross-metathesis of unsaturated α-diazo-β-keto esters using Grubbs’ 2nd generation catalyst, followed by Rh2(OAc)4-catalysed tandem carbonyl ylide formation-intramolecular cycloaddition is demonstrated. The two different catalytic metallocarbene transfer reactions have also been successfully carried out in a one-pot procedure, which allows rapid generation of molecular complexity in a single operation.
Co-reporter:David M. Hodgson Dr.;Steven M. Miles Dr.
Angewandte Chemie 2006 Volume 118(Issue 6) pp:
Publication Date(Web):27 DEC 2005
DOI:10.1002/ange.200503303
Aus zwei mach eins: Die Dimerisierung enantiomerenreiner terminaler Aziridine durch Lithiierung liefert effizient N-geschützte 2-En-1,4-diamine mit vollständiger Selektivität für das E-Olefin (siehe Schema). Die Nützlichkeit der Methode wurde durch die Synthese von (R,S,S,R)-2,5-Diamino-1,6-diphenylhexan-3,4-diol demonstriert, der zentralen Einheit vieler äußerst potenter HIV-Protease-Inhibitoren und asymmetrischer Katalysatoren.
Co-reporter:David M. Hodgson,Steven M. Miles
Angewandte Chemie International Edition 2006 45(6) pp:935-938
Publication Date(Web):
DOI:10.1002/anie.200503303
Co-reporter:David M. Hodgson and Deepshikha Angrish
Chemical Communications 2005 (Issue 39) pp:4902-4904
Publication Date(Web):31 Aug 2005
DOI:10.1039/B510829B
Grubbs' 2nd-generation ruthenium carbene catalyst homocouples diazoacetates to maleates and also catalyses head-to-head dimerisation of alkenyl diazoacetates giving dienyl dilactones.
Co-reporter:David M. Hodgson and Frédéric Le Strat
Chemical Communications 2004 (Issue 7) pp:822-823
Publication Date(Web):24 Feb 2004
DOI:10.1039/B316908A
1,3-Dipolar addition of allene to the carbonyl ylide derived from 6-diazoheptane-2,5-dione is the key step in syntheses of 3-hydroxy-cis-nemorensic acid and nemorensic acid.
Co-reporter:David M. Hodgson, Bogdan Štefane, Timothy J. Miles and Jason Witherington
Chemical Communications 2004 (Issue 19) pp:2234-2235
Publication Date(Web):24 Aug 2004
DOI:10.1039/B409486G
Organolithium-induced ring-opening of aziridines of 2,5-dihydrofuran (5 and 8) and 1,4-dimethoxybut-2-ene (16, 17 and 23) gives 3-substituted 2-aminobut-3-en-1-ols 9–15 and amino ethers 18–20 and 24–26.
Co-reporter:David M. Hodgson;Tobias Brückl;Rebecca Glen;Agnès H. Labande;Deborah A. Selden;Alexander G. Dossetter;Alison J. Redgrave
PNAS 2004 Volume 101 (Issue 15 ) pp:5450-5454
Publication Date(Web):2004-04-13
DOI:10.1073/pnas.0307274101
Catalyzed cascade reactions that generate molecular complexity rapidly and in an enantioselective manner are attractive methods
for asymmetric synthesis. In the present article, chiral rhodium catalysts are shown to effect such a transformation by using
a range of 2-diazo-3,6-diketoesters with bicyclo[2.2.1]alkenes and styrenes as reaction partners. The reactions are likely
to proceed by formation of a catalyst-complexed carbonyl ylide from the diazo compound, followed by intermolecular cycloaddition
with the alkene dipolarophile. It was possible to obtain high levels of asymmetric induction [up to 89% enantiomeric excess
(ee) and 92% ee for the two chiral catalysts investigated]. Enantioselectivity is not highly sensitive to substituent variation
at the ketone that forms the ylide; however, branching does improve ee. Observations of dipolarophile-dependent enantiofacial
selectivity in the cycloadditions indicate that the dipolarophile can be intimately involved in the enantiodiscrimination
process.
Co-reporter:David M. Hodgson, Matthew A. H. Stent, Bogdan Štefane and Francis X. Wilson
Organic & Biomolecular Chemistry 2003 vol. 1(Issue 7) pp:1139-1150
Publication Date(Web):17 Feb 2003
DOI:10.1039/B212404A
A screen of external chiral ligands has led to enantioselective organolithium-induced alkylative double ring-opening of 3,4-epoxytetrahydrofuran 1 with n-BuLi to give 3-methyleneheptane-1,2-diol 3 in 75% yield and 55% ee in the presence of bisoxazoline 10, and in up to 60% ee in the presence of (−)-sparteine 2. Extending the alkylative double ring-opening reaction to epoxides derived from oxabicyclo[n.2.1]alkenes (n
= 2, 3) results in the formation of cycloalkenediols, which, when carried out in the presence of (−)-sparteine 2 affords products in up to 85% ee.
Co-reporter:David M. Hodgson, Magnus W. P. Bebbington and Paul Willis
Organic & Biomolecular Chemistry 2003 vol. 1(Issue 21) pp:3787-3798
Publication Date(Web):22 Aug 2003
DOI:10.1039/B306717N
Radical thiol additions to 7-azanorbornadienes give 7-thio-substituted 2-azanorbornenes and Barton deoxygenations of 7-azabenzonorbornanols give 2-azabenzonorbornanes. The processes both involve novel nitrogen-directed radical rearrangements. The kinetics and mechanisms of the reactions are also discussed.
Co-reporter:David M. Hodgson, Timothy J. Buxton, Iain D. Cameron, Emmanuel Gras and Eirene H. M. Kirton
Organic & Biomolecular Chemistry 2003 vol. 1(Issue 23) pp:4293-4301
Publication Date(Web):24 Oct 2003
DOI:10.1039/B309717J
Enantioselective α-deprotonation of achiral epoxides 1, 21, and 26 using organolithiums in the presence of (−)-sparteine 2 and subsequent electrophile trapping gives access to enantioenriched trisubstituted epoxides 9–17, 22, 23, 27 and 28
(in up to 86% ee).
Co-reporter:David M. Hodgson, Deborah A. Selden, Alexander G. Dossetter
Tetrahedron: Asymmetry 2003 Volume 14(Issue 24) pp:3841-3849
Publication Date(Web):12 December 2003
DOI:10.1016/j.tetasy.2003.09.029
The synthesis of tetrakis[4,4′,6,6′-tetrasubstituted-1,1′-bi-2-naphtholphosphate]dirhodium(II) complexes, and their use as catalysts in the enantioselective tandem carbonyl ylide formation–intramolecular 1,3-dipolar cycloaddition of an unsaturated 2-diazo-3,6-diketoester, generating cycloadduct in up to 86% ee, is described.Graphic(S)-4,4′,6,6′-Tetraoctynyl-1,1′-bi-2-naphtholC52H62O2[α]D=+78.9 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: S(S)-4,4′,6,6′-Tetra(2-phenylethynyl)-1,1′-bi-2-naphtholC52H30O2[α]D=+137.5 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: S(S)-4,4′,6,6′-Tetra(2-phenylethyl)-1,1′-bi-2-naphtholC52H46O2[α]D=+15.0 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: S(S)-4,4′,6,6′-Tetra(4-n-butylphenyl)-1,1′-bi-2-naphtholC60H62O2[α]D=+68.9 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: S(S)-4,4′,6,6′-Tetraoctyl-1,1′-bi-2-naphtholC52H78O2[α]D=+16.3 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: S(S)-4,4′,6,6′-Tetraoctyl-1,1′-bi-2-naphthol phosphateC52H77O4P[α]D=+119 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: S(S)-4,4′,6,6′-Tetra(4-n-butylphenyl)-1,1′-bi-2-naphthol phosphateC60H61O4P[α]D=+82.4 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: S(S)-4,4′,6,6′-Tetraphenyl-1,1′-bi-2-naphthol phosphateC44H29O4P[α]D=+76.5 (c 0.33, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: S(S)-4,4′,6,6′-Tetra(2-phenylethyl)-1,1′-bi-2-naphthol phosphateC52H45O4P[α]D=+104 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: STetrakis[(S)-4,4′,6,6′-tetraoctyl-1,1′-bi-2-naphthol phosphate]dirhodium(II)C208H304O16P4Rh2[α]D=−6.9 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: STetrakis[(S)-4,4′,6,6′-tetra(2-phenylethyl)-1,1′-bi-2-naphthol phosphate]dirhodium(II)C208H176O16P4Rh2[α]D=−14.8 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: STetrakis[(S)-4,4′,6,6′-tetra(4-n-butylphenyl)-1,1′-bi-2-naphtholphosphate]dirhodium(II)C240H240O16P4Rh2[α]D=+219 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: S(1S,5R,7S)-7-Carbo-tert-butoxy-11-oxa-tricyclo[5.3.1.0]undecan-8-oneC15H22O4Ee=86%[α]D=−12.4 (c 1, CHCl3)Source of chirality: RhII (S)-BINOL-derived catalystAbsolute configuration: 1S,5R,7S (from X-ray analysis of (−)-borneol ester)Tetrakis[(S)-4,4′,6,6′-tetraphenyl-1,1′-bi-2-naphthol phosphate]dirhodium(II)C176H112O16P4Rh2[α]D=+215 (c 1, CHCl3)Source of chirality: (S)-BINOLAbsolute configuration: S
Co-reporter:David M. Hodgson, Agnès H. Labande, Rebecca Glen, Alison J. Redgrave
Tetrahedron: Asymmetry 2003 Volume 14(Issue 7) pp:921-924
Publication Date(Web):4 April 2003
DOI:10.1016/S0957-4166(03)00036-3
Catalytic enantioselective tandem carbonyl ylide formation-cycloaddition reactions of tert-butyl 2-diazo-3,6-dioxoheptanoate 7 with alkyne and strained alkene dipolarophiles to afford the corresponding cycloadducts with up to 92% ee are described.Graphic1-tert-Butyloxycarbonyl-5-methyl-7-phenyl-8-oxabicyclo[3.2.1]oct-6-en-2-oneC19H22O4Ee=61% (by chiral HPLC)[α]D24=−171.0 (c 1.0, CHCl3)Source of chirality: Rh (R)-BINOL-derived catalystAbsolute configuration: unknown(1α,4α,4aβ,5α,9α,9aβ)-Decahydro-5-tert-butyloxycarbonyl-9-methyl-5,9-epoxy-1,4-methano-6H-benzocyclohepten-6-oneC18H26O4Ee=92% (by chiral GC)[α]D27=+58.9 (c 1.0, CHCl3)Source of chirality: Rh (R)-BINOL-derived catalystAbsolute configuration: unknown(1α,4α,4aβ,5α,9α,9aβ)-1,4,4a,5,7,8,9,9a-Octahydro-5-tert-butyloxycarbonyl-9-methyl-5,9-epoxy-1,4-methano-6H-benzocyclohepten-6-oneC18H24O4Ee=83% (by chiral GC)[α]D27=+76.2 (c 1.0, CHCl3)Source of chirality: Rh (R)-BINOL-derived catalystAbsolute configuration: unknown
Co-reporter:David M. Hodgson, Jean-Marie Galano and Martin Christlieb
Chemical Communications 2002 (Issue 20) pp:2436-2437
Publication Date(Web):24 Sep 2002
DOI:10.1039/B208485F
The first total synthesis of (−)-xialenon A (1) via conjugate allylation of a 1,5-cyclooctadiene-derived bicyclo[3.3.0]octenone 3 and an α′-hydroxylation on the more hinderd face of enone 9 using hypervalent iodine chemistry, is described.
Co-reporter:David M. Hodgson Dr.;Christopher R. Maxwell;Timothy J. Miles;Edyta Paruch Dr.;Matthew A. H. Stent;Ian R. Matthews Dr.;Francis X. Wilson Dr.;Jason Witherington Dr.
Angewandte Chemie International Edition 2002 Volume 41(Issue 24) pp:
Publication Date(Web):12 DEC 2002
DOI:10.1002/anie.200290003
Co-reporter:David M. Hodgson Dr.;Christopher R. Maxwell;Timothy J. Miles;Edyta Paruch Dr.;Matthew A. H. Stent;Ian R. Matthews Dr.;Francis X. Wilson Dr.;Jason Witherington Dr.
Angewandte Chemie 2002 Volume 114(Issue 24) pp:
Publication Date(Web):12 DEC 2002
DOI:10.1002/ange.200290002
Co-reporter:David M. Hodgson Dr.;Christopher R. Maxwell;Timothy J. Miles;Edyta Paruch Dr.;Matthew A. H. Stent;Ian R. Matthews Dr.;Francis X. Wilson Dr.;Jason Witherington Dr.
Angewandte Chemie 2002 Volume 114(Issue 22) pp:
Publication Date(Web):12 NOV 2002
DOI:10.1002/1521-3757(20021115)114:22<4489::AID-ANGE4489>3.0.CO;2-8
Enantioselektive Erzeugung und intermolekulares Abfangen eines Lithiumcarbenoids erfolgt bei der Reaktion von Epoxiden aus 2,5-Dihydrofuran, 2,5-Dihydropyrrol sowie Oxa- und Aza-bicyclo[n.2.1]alkenen (n=2, 3) mit Organolithiumverbindungen in Gegenwart externer chiraler Liganden. Diese Methode kann zur Synthese wichtiger ungesättigter Diole und Aminoalkohole verwendet werden (siehe Schema; NBoc=N-t-Butoxycarbonyl).
Co-reporter:David M. Hodgson Dr.;Christopher R. Maxwell;Timothy J. Miles;Edyta Paruch Dr.;Matthew A. H. Stent;Ian R. Matthews Dr.;Francis X. Wilson Dr.;Jason Witherington Dr.
Angewandte Chemie International Edition 2002 Volume 41(Issue 22) pp:
Publication Date(Web):12 NOV 2002
DOI:10.1002/1521-3773(20021115)41:22<4313::AID-ANIE4313>3.0.CO;2-B
Enantioselective generation and intermolecular trapping of a lithium carbenoid occurs in the reaction of epoxides of 2,5-dihydrofuran, 2,5-dihydropyrrole, and oxa- and aza-bicyclo[n.2.1]alkenes (n=2, 3) with organolithium compounds in the presence of external chiral ligands. This methodology leads to the synthesis of important unsaturated diol and amino alcohol functionality (see scheme; tert-butoxycarbonyl).
Co-reporter:David M. Hodgson Dr.;Emmanuel Gras Dr.
Angewandte Chemie International Edition 2002 Volume 41(Issue 13) pp:
Publication Date(Web):1 JUL 2002
DOI:10.1002/1521-3773(20020703)41:13<2376::AID-ANIE2376>3.0.CO;2-Q
Simple meso-epoxides can be asymmetrically functionalized: Ligand-assisted direct hydrogen–lithium exchange allows the generation of destabilized oxiranyl lithium species and their subsequent trapping by a wide array of electrophiles (see scheme; E=group formed by addition of electrophile). When carried out in the presence of (−)-sparteine as ligand, this reaction provides a new enantioselective route to epoxides.
Co-reporter:David M. Hodgson, Françoise Y. T. M. Pierard and Paul A. Stupple
Chemical Society Reviews 2001 vol. 30(Issue 1) pp:50-61
Publication Date(Web):27 Nov 2000
DOI:10.1039/B000708K
This review summarizes the background and state of the art (as
of June 2000) of catalytic enantioselective rearrangements and
cycloadditions involving ylides from diazo compounds. The field is still in
its infancy, as until a few years ago it was not considered likely that a
catalyst, having decomposed a diazo compound to form an ylide, would still
be able to exert an influence on subsequent transformations of the ylide.
Recent developments have shown this not to be the case and a new era of
synthetically important ylide transformations in organic chemistry has
begun where appropriate metal–chiral (nonracemic) ligand combinations
are starting to be developed to render these transformations
enantioselective.
Co-reporter:David M. Hodgson, Sarah F. Barker, Laura H. Mace and Julian R. Moran
Chemical Communications 2001 (Issue 2) pp:153-154
Publication Date(Web):21 Dec 2000
DOI:10.1039/B008565K
The isomerisation of vinyldisilanes 1 to allyldisilanes 2
catalysed by palladium-on-carbon in diethyl ether under hydrogen is
described.
Co-reporter:David M. Hodgson, Magnus W. P. Bebbington and Paul Willis
Chemical Communications 2001 (Issue 10) pp:889-890
Publication Date(Web):20 Apr 2001
DOI:10.1039/B101984H
Addition of thiols to 7-azabicyclo[2.2.1]heptadienes such as
16 leads exclusively to 7-thio-substituted 2-azabicyclo[2.2.1]-hept-5-enes
17 in good yields via tandem intermolecular radical
addition—homoallylic radical rearrangement.
Co-reporter:David M. Hodgson, Margarita Petroliagi
Tetrahedron: Asymmetry 2001 Volume 12(Issue 6) pp:877-881
Publication Date(Web):17 April 2001
DOI:10.1016/S0957-4166(01)00147-1
The enantioselective intramolecular oxonium ylide formation-[3,2] sigmatropic rearrangement of α-diazo-β-ketoesters 9 using catalytic (1 mol%) dirhodium tetrakisbinaphthol phosphate catalysts 1 and 2 to give benzofuranones 11 in up to 62% e.e. is described.Graphic2-Allyl-2-methyloxycarbonyl-2,3-dihydrobenzofuran-3-oneC13H12O4Ee=62% (by chiral HPLC)[α]D24=+28.0 (c 0.25, CHCl3)Source of chirality: Rh R-BINOL-derived catalystAbsolute configuration: unknown2-Allyl-2-t-butoxycarbonyl-2,3-dihydrobenzofuran-3-oneC16H18O4Ee=45% (by chiral HPLC)[α]D24=+19.6 (c 0.5, CHCl3)Source of chirality: Rh R-BINOL-derived catalystAbsolute configuration: unknown
Co-reporter:David M. Hodgson Dr.;Paul A. Stupple;Françoise Y. T. M. Pierard Dr.;Agnès H. Labe Dr. and;Craig Johnstone Dr.
Chemistry - A European Journal 2001 Volume 7(Issue 20) pp:
Publication Date(Web):5 OCT 2001
DOI:10.1002/1521-3765(20011015)7:20<4315::AID-CHEM4315>3.0.CO;2-B
The cover picture shows a space-filling representation of a chiral dirhodium catalyst that converts diazo-containing substrates into tricyclic adducts (artwork and three-dimensional modelling by Karl Harrison, University of Oxford). The dirhodium tetrakis(binaphtholphosphate) catalyst (octopus-like, with its eight dodecyl arms providing solubility in hydrocarbon solvents) induces up to 90 % enantiomeric excess in this newly emerging asymmetric cascade process. The evolution of the catalyst for the generation and enantioselective 1,3-dipolar cycloaddition of carbonyl ylides is described in detail by D. M. Hodgson et al. on p. 4465 ff.
Co-reporter:David M. Hodgson Dr.;Paul A. Stupple;Françoise Y. T. M. Pierard Dr.;Agnès H. Labe Dr. and;Craig Johnstone Dr.
Chemistry - A European Journal 2001 Volume 7(Issue 20) pp:
Publication Date(Web):5 OCT 2001
DOI:10.1002/1521-3765(20011015)7:20<4465::AID-CHEM4465>3.0.CO;2-W
Catalytic, enantioselective, tandem carbonyl ylide formation/cycloaddition of 2-diazo-3,6-diketoester 2 with the use of dirhodium tetrakiscarboxylate and tetrakisbinaphtholphosphate catalysts to give the cycloadducts 3 in good yields and up to 90 % ee is described.
Co-reporter:David M. Hodgson and Leonard H. Winning
Organic & Biomolecular Chemistry 2007 - vol. 5(Issue 19) pp:NaN3082-3082
Publication Date(Web):2007/08/02
DOI:10.1039/B707566A
Radical rearrangements are important transformations in organic synthesis. The stabilisation of α-nitrogen radicals is shown to be a useful effect for the control of radical rearrangements and is applied to the synthesis of a variety of azabicyclic frameworks. The utility of this method is illustrated in the synthesis of bioactive targets.
Co-reporter:David M. Hodgson, Carolina Villalonga-Barber, Jonathan M. Goodman and Silvina C. Pellegrinet
Organic & Biomolecular Chemistry 2010 - vol. 8(Issue 17) pp:NaN3984-3984
Publication Date(Web):2010/07/02
DOI:10.1039/C004496B
Reaction of diazodiketoesters 17 and 28 with methyl glyoxylate in the presence of catalytic rhodium(II) acetate generates predominantly the 6,8-dioxabicyclo[3.2.1]octanes 29 and 30, respectively. Acid-catalysed rearrangement of the corresponding alcohol 31 favours, at equilibrium, the 2,8-dioxabicyclo[3.2.1]octane skeleton 33 of the squalestatins–zaragozic acids. Force field calculations on the position of the equilibrium gave misleading results. DFT calculations were correct in suggesting that the energy difference between 31 and 33 should be small, but did not always suggest the right major product. Calculation of the NMR spectra of the similar structures could be used to assign the isomers with a high level of confidence.
Co-reporter:David M. Hodgson, Eric P. A. Talbot and Barry P. Clark
Chemical Communications 2012 - vol. 48(Issue 51) pp:NaN6350-6350
Publication Date(Web):2012/05/22
DOI:10.1039/C2CC32283H
The first synthesis of the bioactive sesquiterpene lactone hydroxyanthecotulide is achieved in 7 steps, involving a stereocontrolled Cr(II)-catalysed reaction of 3-(bromomethyl)furan-2(5H)-one with enynal 9 and a mild Au(I)-catalysed Meyer–Schuster rearrangement.
Co-reporter:David M. Hodgson and Tanzeel Arif
Chemical Communications 2011 - vol. 47(Issue 9) pp:NaN2687-2687
Publication Date(Web):2011/01/12
DOI:10.1039/C0CC04429F
β-Lithiooxyphosphonium ylides, generated in situ from aldehydes and methylenetriphenylphosphorane, react with halomethyl esters to form disubstituted allylic esters in good yields and with high Z-selectivity.
Co-reporter:David M. Hodgson, Elena Moreno-Clavijo, Sophie E. Day and Stanislav Man
Organic & Biomolecular Chemistry 2013 - vol. 11(Issue 32) pp:NaN5369-5369
Publication Date(Web):2013/07/12
DOI:10.1039/C3OB41251B
A stereocontrolled synthesis of norphenyl hyperolactone C together with its utility as a direct precursor to the anti-HIV agent hyperolactone C and analogues by addition of organolithiums, are described. Preliminary studies to access this key building block in a catalytic enantioselective manner are also reported.
Co-reporter:David M. Hodgson and Rosanne S. D. Persaud
Organic & Biomolecular Chemistry 2012 - vol. 10(Issue 39) pp:NaN7951-7951
Publication Date(Web):2012/08/30
DOI:10.1039/C2OB26346G
β-Lithiooxyphosphonium ylides, made in situ from an aldehyde and methylenetriphenylphosphorane, react with a second aldehyde to form E-allylic alcohols. α-Branching and α,β-unsaturation in the second aldehyde, together with the lack of further substitution on the phosphorane carbon play important roles in selectivity. A range of these aldehydes, in addition to aromatic aldehydes as the second aldehyde also provided synthetically useful access to E-allylic alcohols.
Co-reporter:Reece Jacques, Ritashree Pal, Nicholas A. Parker, Claire E. Sear, Peter W. Smith, Aubert Ribaucourt and David M. Hodgson
Organic & Biomolecular Chemistry 2016 - vol. 14(Issue 25) pp:NaN5893-5893
Publication Date(Web):2016/04/25
DOI:10.1039/C6OB00593D
In the past two decades, alkene metathesis has risen in prominence to become a significant synthetic strategy for alkene formation. Many total syntheses of natural products have used this transformation. We review the use, from 2003 to 2015, of ring-closing alkene metathesis (RCM) for the generation of dihydrofurans or -pyrans in natural product synthesis. The strategies used to assemble the RCM precursors and the subsequent use of the newly formed unsaturation will also be highlighted and placed in context.