Search results

Search for "decarboxylative" in Full Text gives 116 result(s) in Beilstein Journal of Organic Chemistry.

Quinone-catalyzed oxidative deformylation: synthesis of imines from amino alcohols

  • Xinyun Liu,
  • Johnny H. Phan,
  • Benjamin J. Haugeberg,
  • Shrikant S. Londhe and
  • Michael D. Clift

Beilstein J. Org. Chem. 2017, 13, 2895–2901, doi:10.3762/bjoc.13.282

Graphical Abstract
  • decarboxylative homologation of α-amino acids [32], which demonstrated for the first time that quinone organocatalysts can be utilized to enable oxidative C–C bond cleavage to provide versatile imine intermediates. To further exploit the utility of this chemistry, we sought to develop a new method for the
  • iminoquinone intermediate is likely required for productive reactivity. To demonstrate the synthetic utility of this methodology, we performed a sequential oxidative deformylation/Mukaiyama−Mannich addition under our previously reported conditions for decarboxylative amino acid homologation (Scheme 4) [32]. In
PDF
Album
Supp Info
Letter
Published 28 Dec 2017

CF3SO2X (X = Na, Cl) as reagents for trifluoromethylation, trifluoromethylsulfenyl-, -sulfinyl- and -sulfonylation. Part 1: Use of CF3SO2Na

  • Hélène Guyon,
  • Hélène Chachignon and
  • Dominique Cahard

Beilstein J. Org. Chem. 2017, 13, 2764–2799, doi:10.3762/bjoc.13.272

Graphical Abstract
  • decarboxylative trifluoromethylation of α,β-unsaturated carboxylic acids in the presence of CF3SO2Na and TBHP [82]. Various (hetero)arenes were compatible with these reaction conditions
PDF
Album
Full Research Paper
Published 19 Dec 2017

A Brønsted base-promoted diastereoselective dimerization of azlactones

  • Danielle L. J. Pinheiro,
  • Gabriel M. F. Batista,
  • Pedro P. de Castro,
  • Leonã S. Flores,
  • Gustavo F. S. Andrade and
  • Giovanni W. Amarante

Beilstein J. Org. Chem. 2017, 13, 2663–2670, doi:10.3762/bjoc.13.264

Graphical Abstract
  • strongly dependent on the base, as well as the substituent at the C2 position [24]. Once again, no comments concerning the stereochemistry were addressed. Thus, we started this study envisioning the trichloromethylation of azlactones through the decarboxylative potassium trichloroacetate (KTCA) in DMSO
PDF
Album
Supp Info
Letter
Published 13 Dec 2017

Regioselective decarboxylative addition of malonic acid and its mono(thio)esters to 4-trifluoromethylpyrimidin-2(1H)-ones

  • Sergii V. Melnykov,
  • Andrii S. Pataman,
  • Yurii V. Dmytriv,
  • Svitlana V. Shishkina,
  • Mykhailo V. Vovk and
  • Volodymyr A. Sukach

Beilstein J. Org. Chem. 2017, 13, 2617–2625, doi:10.3762/bjoc.13.259

Graphical Abstract
  • , 4 Svobody sq, Kharkiv 61122, Ukraine 10.3762/bjoc.13.259 Abstract Background: Due to the high reactivity towards various C-nucleophiles, trifluoromethylketimines are known to be useful reagents for the synthesis of α-trifluoromethylated amine derivatives. However, decarboxylative reactions with
  • , unique heterocyclic ketimines, react with malonic acid under organic base catalysis to regioselectively provide either Michael- or Mannich-type decarboxylative addition products depending on solvent polarity. Malonic mono(thio)esters give exclusively Michael-type products. The two regioisomeric products
  • synthesis of novel isomeric 4(6)-trifluoromethylated 1,2,3,4-tetrahydro- and perhydro-(2-oxopyrimidin-4-yl)acetic acid derivatives. Keywords: ketimines; malonic acid; Michael- and Mannich-type decarboxylative addition; pyrimidin-2(1H)-ones; regioselectivity; trifluoromethyl group; Introduction
PDF
Album
Supp Info
Full Research Paper
Published 07 Dec 2017

1-Imidoalkylphosphonium salts with modulated Cα–P+ bond strength: synthesis and application as new active α-imidoalkylating agents

  • Jakub Adamek,
  • Roman Mazurkiewicz,
  • Anna Węgrzyk and
  • Karol Erfurt

Beilstein J. Org. Chem. 2017, 13, 1446–1455, doi:10.3762/bjoc.13.142

Graphical Abstract
  • . The crucial step in the method included the decarboxylative α-methoxylation of N-phthaloyl- or N-succinylamino acids to the corresponding N-(1-methoxyalkyl)imides, followed by the displacement of the methoxy group by the triarylphosphonium group through melting of the imide derivative with
  • knowledge, attempts at an electrochemical decarboxylative α-methoxylation of 2-imidoalkanecarboxylic acids have been reported only twice in the literature [33][34]. The reactions were carried out in MeOH in the presence of sodium methoxide. Unfortunately, because of the low reaction selectivity related to
  • side reactions (for example Kolbe-dimerization), the yields were only poor (10–35%) [33][34]. According to our previously reported procedure for the electrochemical decarboxylative α-methoxylation of N-acyl-α-amino acids [18], amino acid derivatives 6 were converted to N-(1-methoxyalkyl)imides 7. The
PDF
Album
Supp Info
Full Research Paper
Published 24 Jul 2017

Sustainable synthesis of 3-substituted phthalides via a catalytic one-pot cascade strategy from 2-formylbenzoic acid with β-keto acids in glycerol

  • Lina Jia and
  • Fuzhong Han

Beilstein J. Org. Chem. 2017, 13, 1425–1429, doi:10.3762/bjoc.13.139

Graphical Abstract
  • demand and potential for the development of a green one-pot cascade aldol/cyclization strategy for these compounds [22]. The use of β-keto acids as ketone enolate equivalents in metal- and organocatalytic decarboxylative aldol reactions has been extensively studied and proven to be a valuable and
  • straightforward method for the preparation of several biologically active compounds of medicinal and agrochemical interest [23][24][25][26][27][28][29][30]. Notably, the decarboxylative reaction of β-keto acids provides a traceless means of activation with CO2 as the only byproduct. On the other hand, glycerol
PDF
Album
Supp Info
Full Research Paper
Published 19 Jul 2017

Decarboxylative and dehydrative coupling of dienoic acids and pentadienyl alcohols to form 1,3,6,8-tetraenes

  • Ghina’a I. Abu Deiab,
  • Mohammed H. Al-Huniti,
  • I. F. Dempsey Hyatt,
  • Emma E. Nagy,
  • Kristen E. Gettys,
  • Sommayah S. Sayed,
  • Christine M. Joliat,
  • Paige E. Daniel,
  • Rupa M. Vummalaneni,
  • Andrew T. Morehead Jr,
  • Andrew L. Sargent and
  • Mitchell P. Croatt

Beilstein J. Org. Chem. 2017, 13, 384–392, doi:10.3762/bjoc.13.41

Graphical Abstract
  • at Greensboro, Greensboro, NC 27402, USA Department of Chemistry, East Carolina University, Greenville, NC 27858, USA 10.3762/bjoc.13.41 Abstract Dienoic acids and pentadienyl alcohols are coupled in a decarboxylative and dehydrative manner at ambient temperature using Pd(0) catalysis to generate
  • 1,3,6,8-tetraenes. Contrary to related decarboxylative coupling reactions, an anion-stabilizing group is not required adjacent to the carboxyl group. Of mechanistic importance, it appears that both the diene of the acid and the diene of the alcohol are required for this reaction. To further understand
  • materials [10][11][12][13][14][15][16][17][18]. Another approach to the formation of C–C bonds is through decarboxylative coupling reactions (Scheme 1). This can be arrived in a one-component fashion via the removal of CO2 from an ester or in a two-component manner by removal of CO2 from a carboxylic acid
PDF
Album
Supp Info
Full Research Paper
Published 28 Feb 2017

Polyketide stereocontrol: a study in chemical biology

  • Kira J. Weissman

Beilstein J. Org. Chem. 2017, 13, 348–371, doi:10.3762/bjoc.13.39

Graphical Abstract
  • appropriate precursor from the cellular pool, a ketosynthase (KS) which extends the chain via a Claisen-like decarboxylative condensation, and a non-catalytic acyl carrier protein (ACP) to which the intermediates are covalently tethered through a phosphopantetheine prosthetic group. The modules can also
PDF
Album
Review
Published 24 Feb 2017

Multicomponent synthesis of spiropyrrolidine analogues derived from vinylindole/indazole by a 1,3-dipolar cycloaddition reaction

  • Manjunatha Narayanarao,
  • Lokesh Koodlur,
  • Vijayakumar G. Revanasiddappa,
  • Subramanya Gopal and
  • Susmita Kamila

Beilstein J. Org. Chem. 2016, 12, 2893–2897, doi:10.3762/bjoc.12.288

Graphical Abstract
  • out by reacting N-alkylvinyl products 3 with azomethine ylide, generated in situ through decarboxylative condensation of ninhydrin (4) and sarcosine (5). The 1,3-dipolar cycloaddition of the ylide with the olefin 3 yielded spiropyrrolidines 7 with regiospecificity (Table 1, entries 1–4). The formation
PDF
Album
Supp Info
Full Research Paper
Published 29 Dec 2016

Chemical probes for competitive profiling of the quorum sensing signal synthase PqsD of Pseudomonas aeruginosa

  • Michaela Prothiwa,
  • Dávid Szamosvári,
  • Sandra Glasmacher and
  • Thomas Böttcher

Beilstein J. Org. Chem. 2016, 12, 2784–2792, doi:10.3762/bjoc.12.277

Graphical Abstract
  • ]. The biosynthesis of AQs has been matter of a long-standing debate that could only recently be resolved. Although HHQ could be produced in vitro by a PqsD catalyzed “head-to-head” decarboxylative Claisen condensation of activated anthranilic acid with β-keto fatty acid derivatives [10][11], isotope
  • catalyzes the condensation with malonyl-CoA to form 2-aminobenzoylacetyl-CoA. The thioesterase PqsE hydrolyses the thioester to produce 2-aminobenzoylacetate (2-ABA) [13]. The PqsBC complex finally generates HHQ or other AQs in a decarboxylative condensation reaction of 2-ABA with fatty acids loaded on PqsC
PDF
Album
Supp Info
Full Research Paper
Published 20 Dec 2016

Enantioselective addition of diphenyl phosphonate to ketimines derived from isatins catalyzed by binaphthyl-modified organocatalysts

  • Hee Seung Jang,
  • Yubin Kim and
  • Dae Young Kim

Beilstein J. Org. Chem. 2016, 12, 1551–1556, doi:10.3762/bjoc.12.149

Graphical Abstract
  • of organocatalysts [38][39][40][41][42][43][44][45], we have reported the catalytic asymmetric decarboxylative aldol addition reaction of isatins with benzoylacetic acids catalyzed by chiral binaphthyl-based squaramide [46]. Here we wish to report the enantioselective addition reaction of diphenyl
PDF
Album
Supp Info
Letter
Published 20 Jul 2016

Biosynthesis of oxygen and nitrogen-containing heterocycles in polyketides

  • Franziska Hemmerling and
  • Frank Hahn

Beilstein J. Org. Chem. 2016, 12, 1512–1550, doi:10.3762/bjoc.12.148

Graphical Abstract
  • -hydroxytetrahydro-2H-pyran-2-one 124 is formally dehydrated by consecutive malonylation–elimination to finally give a 5,6-dihydro-2H-pyran-2-one 127 [127]. The tailoring enzyme PlmT2 was proposed to catalyse the decarboxylative elimination of malonoyl halfester 126. It is not clear, whether the initial malonylation
PDF
Album
Review
Published 20 Jul 2016

Multicomponent reactions: A simple and efficient route to heterocyclic phosphonates

  • Mohammad Haji

Beilstein J. Org. Chem. 2016, 12, 1269–1301, doi:10.3762/bjoc.12.121

Graphical Abstract
  • ]. The desired phosphono-substituted pyrroles were isolated in 41–87% yield under solvent and catalyst-free conditions. Kaboudin et al. described a three-component, catalyst-free decarboxylative coupling of proline (296) with aldehydes 297 and dialkyl phosphonates to afford pyrrolidinylphosphonates 300
  • . Multicomponent reaction of alkanedials, acetamide and acetyl chloride in the presence of PCl3 and acetic acid. An oxidative domino three-component synthesis of polyfunctionalized pyridines. A sequential one-pot three-component synthesis of polysubstituted pyrroles. Three-component decarboxylative coupling of
PDF
Album
Review
Published 21 Jun 2016

Synthesis of 2-oxindoles via 'transition-metal-free' intramolecular dehydrogenative coupling (IDC) of sp2 C–H and sp3 C–H bonds

  • Nivesh Kumar,
  • Santanu Ghosh,
  • Subhajit Bhunia and
  • Alakesh Bisai

Beilstein J. Org. Chem. 2016, 12, 1153–1169, doi:10.3762/bjoc.12.111

Graphical Abstract
  • decarboxylative protonation on 2-oxindoles bearing an benzylester or para-methoxybenzyl ester at the 3-position in presence of a catalytic amount of Pd on activated charcoal. We have also shown the direct installation of allyl, prenyl, reverse-prenyl, or geranyl groups at the 3-position of 2-oxindole using Pd
  • -catalyzed decarboxylative strategies [47]. Results and Discussion We decided to use iodine as an oxidant for the synthesis of 2-oxindoles [48][49][50][51][52][53], starting from β-N-arylamido ester 3a and methyl iodide as the substrates (Table 1). An elaborate optimization study suggested that the
  • reaction in the presence of iodine or NIS. We envisioned that the oxidative coupling products containing benzyl or p-methoxybenzyl ester could be effective intermediates for the synthesis of 3-monosubstituted 2-oxindoles via deprotection of the benzyl group followed by decarboxylative protonation in
PDF
Album
Supp Info
Full Research Paper
Published 08 Jun 2016

Catalytic asymmetric synthesis of biologically important 3-hydroxyoxindoles: an update

  • Bin Yu,
  • Hui Xing,
  • De-Quan Yu and
  • Hong-Min Liu

Beilstein J. Org. Chem. 2016, 12, 1000–1039, doi:10.3762/bjoc.12.98

Graphical Abstract
  • catalysts have also been used for the synthesis of chiral 3-hydroxyoxindoles. In 2013, Pan and co-workers reported the Yb(OTf)3-catalyzed enantioselective decarboxylative addition of β-ketoacids to isatins, forming the 3-hydroxyoxindoles in excellent yields (up to 98% yield) and with high
  • isatin derivatives were tolerated under these conditions. The aldol reaction of N-benzyl-5-bromoisatin with 1-thiacyclohexan-4-one and cyclohexanone gave the corresponding products with 96% and 99% ee, respectively. More recently, Kesavan and co-workers reported an asymmetric decarboxylative
  • crucial for the reactivity by forming hydrogen-bond interactions with CPA. Ma and co-workers described the CPA (cat. 32)-catalyzed enantioselective decarboxylative alkylation of β-keto acids with 3-hydroxy-3-indolyloxindoles, affording the 3-functionlized 3-indolyloxindoles bearing an all-carbon
PDF
Album
Review
Published 18 May 2016

Cascade alkylarylation of substituted N-allylbenzamides for the construction of dihydroisoquinolin-1(2H)-ones and isoquinoline-1,3(2H,4H)-diones

  • Ping Qian,
  • Bingnan Du,
  • Wei Jiao,
  • Haibo Mei,
  • Jianlin Han and
  • Yi Pan

Beilstein J. Org. Chem. 2016, 12, 301–308, doi:10.3762/bjoc.12.32

Graphical Abstract
  • [18][19], decarboxylative alkenylation of cycloalkanes with aryl vinylic carboxylic acids [20][21], trifluoromethylthiolation [22], thiolation [23][24], alkenylation [25][26], dehydrogenation−olefination and esterification [27][28], radical addition/1,2-aryl migration [29], cascade alkylation
PDF
Album
Supp Info
Full Research Paper
Published 17 Feb 2016

A convergent, umpoled synthesis of 2-(1-amidoalkyl)pyridines

  • Tarn C. Johnson and
  • Stephen P. Marsden

Beilstein J. Org. Chem. 2016, 12, 1–4, doi:10.3762/bjoc.12.1

Graphical Abstract
  • substitution of suitably-activated pyridine N-oxides by azlactone nucleophiles, followed by decarboxylative azlactone ring-opening. The synthesis obviates the need for precious metal catalysts to achieve a formal enolate arylation reaction, and constitutes a formally ‘umpoled’ approach to this valuable class
  • arylation/decarboxylative hydrolysis approach to 2-(1-amidoalkyl)pyridines. Substrate scope of the direct amidoalkylation of pyridine N-oxides. Supporting Information Supporting Information File 7: Experimental procedures and full compound characterisation data for products 8a–j. Supporting Information
PDF
Album
Supp Info
Letter
Published 04 Jan 2016

Carbon–carbon bond cleavage for Cu-mediated aromatic trifluoromethylations and pentafluoroethylations

  • Tsuyuka Sugiishi,
  • Hideki Amii,
  • Kohsuke Aikawa and
  • Koichi Mikami

Beilstein J. Org. Chem. 2015, 11, 2661–2670, doi:10.3762/bjoc.11.286

Graphical Abstract
  • the decarboxylative trifluoromethylation of aryl halides [37] (Scheme 4). Not only iodobenzene but also 4-bromotoluene was trifluoromethylated by the [(NHC)Cu(TFA)] complex. The perfluoroalkylation reactions mentioned above require a stoichiometric amount of copper reagent, whereas it was found that
  • the addition of silver salts is effective for the copper-mediated trifluoromethylation of aryl iodides [38] (Scheme 5). The amount of copper used in the reaction was reduced to 30 or 40 mol % by adding a small amount of Ag2O. As a related decarboxylative transformation, silver-mediated aromatic
  • trifluoromethylation of aryl iodides with ClCF2CO2Me and fluoride can be utilized for clinical studies. Herein, we introduce one example of decarboxylative [18F]trifluoromethylation for positron emission tomography (PET) studies. A synthetic methodology for [18F]labelled-CF3 arenes is desired for the application of
PDF
Album
Review
Published 18 Dec 2015

Preparative semiconductor photoredox catalysis: An emerging theme in organic synthesis

  • David W. Manley and
  • John C. Walton

Beilstein J. Org. Chem. 2015, 11, 1570–1582, doi:10.3762/bjoc.11.173

Graphical Abstract
  • , cyclic amines were produced. Carboxylic acids were particularly fruitful affording C-centered radicals that alkylated alkenes and took part in tandem addition cyclizations producing chromenopyrroles; decarboxylative homo-dimerizations were also observed. Acceptors initially yielding radical anions
  • components, decarboxylative arylations of amino acids, diastereoselective preparations of cis-cyclobutanes via [2 + 2] cycloadditions of enones, selective reductions of benzylic and α-carbonyl halides and, with fac-Ir(ppy)3, reductions of unactivated alkyl iodides [5][6][7][8]. Furthermore, the value of
  • functionalized carboxylic acids that take part in decarboxylative additions to maleimides is presented in Scheme 5. Reasonable yields of adducts were obtained for alkyl radicals with α-alkoxy substituents. Two diastereomers of 22 as a 1:1 mixture were obtained from 2-tetrahydrofuroic acid in a very pleasing
PDF
Album
Review
Published 09 Sep 2015

Radical-mediated dehydrative preparation of cyclic imides using (NH4)2S2O8–DMSO: application to the synthesis of vernakalant

  • Dnyaneshwar N. Garad,
  • Subhash D. Tanpure and
  • Santosh B. Mhaske

Beilstein J. Org. Chem. 2015, 11, 1008–1016, doi:10.3762/bjoc.11.113

Graphical Abstract
  • persistent interest [28][29][30][31][32][33][34][35][36][37][38][39][40][41][42][43][44][45][46][47][48][49][50][51][52][53][54]. Results and Discussion While working on the development of a palladium-catalyzed decarboxylative C–H activation methodology to access the important core structure dihydroquinolone
PDF
Album
Supp Info
Full Research Paper
Published 12 Jun 2015

Sequential decarboxylative azide–alkyne cycloaddition and dehydrogenative coupling reactions: one-pot synthesis of polycyclic fused triazoles

  • Kuppusamy Bharathimohan,
  • Thanasekaran Ponpandian,
  • A. Jafar Ahamed and
  • Nattamai Bhuvanesh

Beilstein J. Org. Chem. 2014, 10, 3031–3037, doi:10.3762/bjoc.10.321

Graphical Abstract
  • describe a one-pot protocol for the synthesis of a novel series of polycyclic triazole derivatives. Transition metal-catalyzed decarboxylative CuAAC and dehydrogenative cross coupling reactions are combined in a single flask and achieved good yields of the respective triazoles (up to 97% yield). This
  • methodology is more convenient to produce the complex polycyclic molecules in a simple way. Keywords: copper(II) acetate; decarboxylative CuAAC; dehydrogenative coupling; fused triazoles; one-pot synthesis; Introduction The copper-catalyzed Huisgen [3 + 2] cycloaddition (or copper-catalyzed azide–alkyne
  • cycloaddition, CuAAC) between an organic azide and a terminal alkyne is a well-established strategy for the construction of 1,4-disubstituted 1,2,3-triazoles [1][2][3][4]. In a recent development, this decarboxylative coupling reaction was well documented for the generation of C–C bonds [5]. This method has
PDF
Album
Supp Info
Letter
Published 17 Dec 2014

Formal total syntheses of classic natural product target molecules via palladium-catalyzed enantioselective alkylation

  • Yiyang Liu,
  • Marc Liniger,
  • Ryan M. McFadden,
  • Jenny L. Roizen,
  • Jacquie Malette,
  • Corey M. Reeves,
  • Douglas C. Behenna,
  • Masaki Seto,
  • Jimin Kim,
  • Justin T. Mohr,
  • Scott C. Virgil and
  • Brian M. Stoltz

Beilstein J. Org. Chem. 2014, 10, 2501–2512, doi:10.3762/bjoc.10.261

Graphical Abstract
  • install the initial stereocenters (Scheme 3). Treatment of 16 with LiHMDS in THF, followed by allyl chloroformate, furnished the known carbonate 17 in high yield [34]. This substrate smoothly undergoes palladium-catalyzed enantioselective decarboxylative allylation in the presence of (S)-t-Bu-PHOX (5
  • enantioenriched piperidinone 47, and thus a single enantiomer of rhazinilam may be prepared. The formal synthesis of (+)-rhazinilam commenced with palladium-catalyzed decarboxylative allylic alkylation of known carboxy-lactam 49 to afford benzoyl-protected piperidinone 50 in 97% yield and 99% ee (Scheme 11) [84
PDF
Album
Supp Info
Full Research Paper
Published 28 Oct 2014

Facile synthesis of 1-alkoxy-1H-benzo- and 7-azabenzotriazoles from peptide coupling agents, mechanistic studies, and synthetic applications

  • Mahesh K. Lakshman,
  • Manish K. Singh,
  • Mukesh Kumar,
  • Raghu Ram Chamala,
  • Vijayender R. Yedulla,
  • Domenick Wagner,
  • Evan Leung,
  • Lijia Yang,
  • Asha Matin and
  • Sadia Ahmad

Beilstein J. Org. Chem. 2014, 10, 1919–1932, doi:10.3762/bjoc.10.200

Graphical Abstract
  • derivative under palladium-catalyzed conditions. The benzoyl ester of BtOH has been evaluated in a decarboxylative Pd-mediated Heck reaction, leading to a modest product yield [36]. However, this appears to be the only example of a BtOH derivative in Pd-mediated reactions. In principle, formation of π–allyl
PDF
Album
Supp Info
Full Research Paper
Published 19 Aug 2014

Multicomponent reactions in nucleoside chemistry

  • Mariola Koszytkowska-Stawińska and
  • Włodzimierz Buchowicz

Beilstein J. Org. Chem. 2014, 10, 1706–1732, doi:10.3762/bjoc.10.179

Graphical Abstract
  • rationalized using semi-empirical calculations. In the same contribution, the cascade reactions starting from uracil polyoxin C 106 were described (Scheme 44). Decarboxylative formation of azomethine ylides from 106 and an aldehyde (or ketone), followed by reaction of the ylide with maleimide afforded mixtures
PDF
Album
Review
Published 29 Jul 2014

A convenient enantioselective decarboxylative aldol reaction to access chiral α-hydroxy esters using β-keto acids

  • Zhiqiang Duan,
  • Jianlin Han,
  • Ping Qian,
  • Zirui Zhang,
  • Yi Wang and
  • Yi Pan

Beilstein J. Org. Chem. 2014, 10, 969–974, doi:10.3762/bjoc.10.95

Graphical Abstract
  • University, Nanjing, 210093, China 10.3762/bjoc.10.95 Abstract We show a convenient decarboxylative aldol process using a scandium catalyst and a PYBOX ligand to generate a series of highly functionalized chiral α-hydroxy esters. The protocol tolerates a broad range of β-keto acids with inactivated aromatic
  • appears as a common procedure, affording chiral tertiary alcohols which are ubiquitous in the biological sciences and pharmaceutical industry [1][2][3][4][5][6]. The decarboxylative aldol reaction, broadly used for the generation of ester enolate equivalents by the promotion of releasing CO2, has become
  • an appealing method to access chiral tertiary alcohols. Taking advantage of this rigid reactivity, several unique catalytic decarboxylative aldol transformations of β-keto acids with various protic aldehydes have been developed [7][8][9][10] (Figure 1). High enantioselectivities were achieved with
PDF
Album
Supp Info
Full Research Paper
Published 29 Apr 2014
Other Beilstein-Institut Open Science Activities