Lukas Hintermann

Find an error

Name:
Organization: Technische Universit?t München , Germany
Department: Department Chemie
Title: Professor(PhD)

TOPICS

Co-reporter:Lukas Hintermann;Kit Ming Wong
European Journal of Organic Chemistry 2017 Volume 2017(Issue 37) pp:5527-5536
Publication Date(Web):2017/10/10
DOI:10.1002/ejoc.201700677
The enantiopure reagent menthyl chloride (2) is generally prepared from (–)-(1R)-menthol (1) with Lucas' reagent (ZnCl2 in conc. aqueous HCl) in a stereoretentive reaction that appeared to be free from accompanying rearrangements. The same was assumed for a recent synthesis of 2 through TiCl4-catalyzed extrusion of SO2 from menthyl chlorosulfite (3). The products of both syntheses have now been analyzed by quantitative 1H and 13C NMR methods, and all reaction components have been identified down to the ≤ 0.5 mol-% level. Either reaction is accompanied by cationic rearrangement to the considerable extent of 18–25 mol-%. Besides the expected 2, neomenthylchloride (4) and five rearrangement products have been identified, among them three regioisomeric tertiary chloromenthanes (9, 10, 11), and both a secondary (12) and tertiary chloride (16) derived from ψ-menthane (1-isobutyl-3-methylcyclopentane). A scheme of rearrangement pathways starting from a common menthyl carbenium ion pair is derived. The effect of purification protocols on crude 2 has been studied quantitatively. Either selective solvolysis of tertiary sideproducts (98 mol-% purity) or low-temperature crystallization (≥ 97 mol-% purity) was successful. An improved, scalable synthesis of 2 through the catalytic rearrangement of chlorosulfite 3 is reported.
Co-reporter:Dr. Andreas Brunner ;Dr. Lukas Hintermann
Chemistry - A European Journal 2016 Volume 22( Issue 8) pp:2787-2792
Publication Date(Web):
DOI:10.1002/chem.201504248

Abstract

Terminal alkynes (RCCH) are homologated by a sequence of ruthenium-catalyzed anti-Markovnikov hydration of alkyne to aldehyde (RCH2CHO), followed by Bestmann–Ohira alkynylation of aldehyde to chain-elongated alkyne (RCH2CCH). Inverting the sequence by starting from aldehyde brings about the reciprocal homologation of aldehydes instead. The use of 13C-labeled Bestmann–Ohira reagent (dimethyl ((1-13C)-1-diazo-2-oxopropyl)phosphonate) for alkynylation provides straightforward access to singly or, through additional homologation, multiply 13C-labeled alkynes. The labeled alkynes serve as synthetic platform for accessing a multitude of specifically 13C-labeled products. Terminal alkynes with one or two 13C-labels in the alkyne unit have been submitted to alkyne–azide click reactions; the copper-catalyzed version (CuAAC) was found to display a regioselectivity of >50 000:1 for the 1,4- over the 1,5-triazine isomer, as shown analytically by 13C NMR spectroscopy.

Co-reporter:M.Sc. Alois Bräuer;Dr. Philipp Beck;Dr. Lukas Hintermann;Dr. Michael Groll
Angewandte Chemie International Edition 2016 Volume 55( Issue 1) pp:422-426
Publication Date(Web):
DOI:10.1002/anie.201507835

Abstract

Multienzymatic cascades are responsible for the biosynthesis of natural products and represent a source of inspiration for synthetic chemists. The FeII/α-ketoglutarate-dependent dioxygenase AsqJ from Aspergillus nidulans is outstanding because it stereoselectively catalyzes both a ferryl-induced desaturation reaction and epoxidation on a benzodiazepinedione. Interestingly, the enzymatically formed spiro epoxide spring-loads the 6,7-bicyclic skeleton for non-enzymatic rearrangement into the 6,6-bicyclic scaffold of the quinolone alkaloid 4′-methoxyviridicatin. Herein, we report different crystal structures of the protein in the absence and presence of synthesized substrates, surrogates, and intermediates that mimic the various stages of the reaction cycle of this exceptional dioxygenase.

Co-reporter:M.Sc. Alois Bräuer;Dr. Philipp Beck;Dr. Lukas Hintermann;Dr. Michael Groll
Angewandte Chemie 2016 Volume 128( Issue 1) pp:432-436
Publication Date(Web):
DOI:10.1002/ange.201507835

Abstract

Für die Biosynthese von Naturstoffen sind Multi-Enzymkaskaden verantwortlich, die generell eine Inspirationsquelle für Synthesechemiker darstellen. Die FeII/α-Ketoglutarat-abhängige Dioxygenase AsqJ aus Aspergillus nidulans ist außergewöhnlich, da sie stereoselektiv sowohl eine Desaturierung als auch Epoxidierung eines Benzodiazepindions katalysiert. Interessanterweise löst das enzymatisch gebildete, gespannte Spiro-Epoxid die enzymunabhängige Umlagerung des 6,7-Bicyclus zum 6,6-Chinolongerüst des Alkaloids 4′-Methoxyviridicatin aus. Verschiedene Kristallstrukturen des Proteins in Ab- und Anwesenheit synthetisierter Substrate, Substratanaloga und Zwischenprodukte erlaubten uns, die einzelnen Stufen der Reaktionssequenz dieser einzigartigen Dioxygenase nachzuahmen und zu hinterleuchten.

Co-reporter:M.Sc. Johannes Schlüter;M.Sc. Max Blazejak;Dr. Florian Boeck ;Dr. Lukas Hintermann
Angewandte Chemie International Edition 2015 Volume 54( Issue 13) pp:4014-4017
Publication Date(Web):
DOI:10.1002/anie.201409252

Abstract

The asymmetric catalytic addition of alcohols (phenols) to non-activated alkenes has been realized through the cycloisomerization of 2-allylphenols to 2-methyl-2,3-dihydrobenzofurans (2-methylcoumarans). The reaction was catalyzed by a chiral titanium–carboxylate complex at uncommonly high temperatures for asymmetric catalytic reactions. The catalyst was generated by mixing titanium isopropoxide, the chiral ligand (aS)-1-(2-methoxy-1-naphthyl)-2-naphthoic acid or its derivatives, and a co-catalytic amount of water in a ratio of 1:1:1 (5 mol % each). This homogeneous thermal catalysis (HOT-CAT) gave various (S)-2-methylcoumarans with yields of up to 90 % and in up to 85 % ee at 240 °C, and in 87 % ee at 220 °C.

Co-reporter:Li Xiao;Alexander Pöthig
Monatshefte für Chemie - Chemical Monthly 2015 Volume 146( Issue 9) pp:1529-1539
Publication Date(Web):2015 September
DOI:10.1007/s00706-015-1515-7
The chemistry of 4,6-dialkyl-2-amino-1,3,5-triazines with bulky alkyl substituents was investigated and their use as building blocks for preparing chiral thiourea organocatalysts explored. Reaction of ammonia with 4,6-di-tert-butyl-2-chloro-1,3,5-triazine gave 4,6-di-tert-butyl-1,3,5-triazin-2-amine which formed extended hydrogen-bond networks in the solid state according to X-ray crystallography. Selected heterocyclic amines were converted to isothiocyanates, and the latter reacted with (S,S)-2-(dimethylamino)cyclohexylamine to give enantiopure 1-hetaryl-3-[2-(dimethylamino)cyclohexyl]thioureas, with hetaryl representing either 4,6-dimethyl-1,3-diazin-2-yl, 4,6-diisopropyl-1,3,5-triazin-2-yl, or 4,6-di-tert-butyl-1,3,5-triazin-2-yl groups. These compounds are structural analogs of Takemotos’s chiral thiourea organocatalysts (1-[3,5-bis(trifluoromethyl)phenyl]-3-[(1S,2S)-2-(dimethylamino)cyclohexyl]thiourea) with an aza-aryl instead of the 3,5-bis(trifluoromethyl)phenyl group. They feature a strong intramolecular N–H to N-1 hydrogen bond, as shown by X-ray crystallography of 1-(4,6-di-tert-butyl-1,3,5-triazin-2-yl)-3-[2-(dimethylamino)cyclohexyl]thiourea in the solid state and by 1H NMR spectroscopy of all derivatives in CDCl3 solution, which prevents them from acting as bifunctional organocatalyst. In the reaction of 4,6-di-tert-butyl-2-chloro-1,3,5-triazine with ammonia, 4,6-di-tert-butyl-2-ethoxy-1,3,5-triazine was identified as side-product displaying a mildly sweet, floral odor that is unusual for a 1,3,5-triazine. Analogs (>35) of 4,6-di-tert-butyl-2-ethoxy-1,3,5-triazine were prepared to define the important structural factors of the olfactophore.
Co-reporter:M.Sc. Johannes Schlüter;M.Sc. Max Blazejak;Dr. Florian Boeck ;Dr. Lukas Hintermann
Angewandte Chemie 2015 Volume 127( Issue 13) pp:4086-4089
Publication Date(Web):
DOI:10.1002/ange.201409252

Abstract

The asymmetric catalytic addition of alcohols (phenols) to non-activated alkenes has been realized through the cycloisomerization of 2-allylphenols to 2-methyl-2,3-dihydrobenzofurans (2-methylcoumarans). The reaction was catalyzed by a chiral titanium–carboxylate complex at uncommonly high temperatures for asymmetric catalytic reactions. The catalyst was generated by mixing titanium isopropoxide, the chiral ligand (aS)-1-(2-methoxy-1-naphthyl)-2-naphthoic acid or its derivatives, and a co-catalytic amount of water in a ratio of 1:1:1 (5 mol % each). This homogeneous thermal catalysis (HOT-CAT) gave various (S)-2-methylcoumarans with yields of up to 90 % and in up to 85 % ee at 240 °C, and in 87 % ee at 220 °C.

Co-reporter:Oleg V. Maltsev, Alexander Pöthig, and Lukas Hintermann
Organic Letters 2014 Volume 16(Issue 5) pp:1282-1285
Publication Date(Web):February 13, 2014
DOI:10.1021/ol500189s
Palladium-catalyzed dehydrosulfurative Liebeskind–Srogl coupling of terminal alkynes with 2-mercapto-1,3-pyrimidine-5-carbaldehyde under base-free conditions provides 2-(alkynyl)-1,3-pyrimidine-5-carbaldehydes, which are substrates for autocatalytic amplification of chirality according to Soai et al. The mercapto aldehyde acceptor is obtained by condensation of Arnold’s vinamidinium salt with thiourea.
Co-reporter:Oleg V. Maltsev;Rodger Rausch;Zheng-Jun Quan
European Journal of Organic Chemistry 2014 Volume 2014( Issue 33) pp:7426-7432
Publication Date(Web):
DOI:10.1002/ejoc.201403133

Abstract

A two-step synthesis of 2-arylpyrimidine-5-carbaldehydes, which are of relevance as substrates for Soai's asymmetric autocatalysis, was realized by exploiting a hidden threefold symmetry in the target core structure. Condensation of Arnold's C3-symmetric vinamidinium cation with S-methylisothiouronium sulfate provides 2-methylsulfanyl-pyrimidine-5-carbaldehyde; introduction of aryl groups at C-2 of the latter was accomplished by a Liebeskind–Srogl palladium-catalyzed desulfurative (de-methylsulfanylative) coupling with aryl boronic acids to obtain the target compounds 1 (14 examples, 60–95 % yield).

Co-reporter:Dr. Oleg V. Maltsev;Dr. Naba K. Nath;Dr. Pan&x10d;e Naumov;Dr. Lukas Hintermann
Angewandte Chemie International Edition 2014 Volume 53( Issue 3) pp:847-850
Publication Date(Web):
DOI:10.1002/anie.201307972

Abstract

The chemistry of firefly bioluminescence is important for numerous applications in biochemistry and analytical chemistry. The emitter of this bioluminescent system, firefly oxyluciferin, is difficult to handle. The cause of its lability was clarified while its synthesis was reinvestigated. A side product was identified and characterized by NMR spectroscopy and X-ray crystallography. The reason for the lability of oxyluciferin is now ascribed to autodimerization of the coexisting enol and keto forms in a Mannich-type reaction.

Co-reporter:Mateusz Rebarz, Boris-Marko Kukovec, Oleg V. Maltsev, Cyril Ruckebusch, Lukas Hintermann, Panče Naumov and Michel Sliwa  
Chemical Science 2013 vol. 4(Issue 10) pp:3803-3809
Publication Date(Web):28 May 2013
DOI:10.1039/C3SC50715G
The mysterious flashes of light communicated by fireflies conceal a rich and exciting solution spectrochemistry that revolves around the chemiexcitation and photodecay of the fluorophore, oxyluciferin. A triple chemical equilibrium by double deprotonation and keto–enol tautomerism turns this simple molecule into an intricate case where the relative spectral contributions of six chemical species combine over a physiologically relevant pH range, rendering physical isolation and spectral characterization of most of the species unmanageable. To disentangle the individual spectral contributors, here we demonstrate the advantage of chemical oriented multivariate data analysis. We designed a set of specific oxyluciferin derivatives and applied a multivariate curve resolution-alternating least squares (MCR-ALS) procedure simultaneously to an extensive set of pH-dependent spectroscopic data for oxyluciferin and the target derivatives. The analysis provided, for the first time, the spectra of the pure individual components free of contributions from the other forms, their pH-dependent profiles and distributions, and the most accurate to date values for the three equilibrium constants.
Co-reporter:Johannes Schlüter;Max Blazejak ;Dr. Lukas Hintermann
ChemCatChem 2013 Volume 5( Issue 11) pp:3309-3315
Publication Date(Web):
DOI:10.1002/cctc.201300182

Abstract

An aluminum-catalyzed intramolecular hydroalkoxylation of nonactivated alkenes is presented as a powerful synthetic tool for the preparation of oxygen heterocycles, which are of major interest for the preparation of biological and pharmaceutical active compounds. The aluminum isopropoxide catalyzed (5 mol %) cyclization of 2-allylphenols at elevated temperatures (250 °C, 20 min) provides 2-methylcoumarans (2-methyl-2,3-dihydrobenzofuran) in an exceptionally fast, simple, and economic manner. Moreover, heating of allyl aryl ethers with aluminum isopropoxide (5 mol %) gives 2-methylcoumarans by a tandem Claisen rearrangement–hydroalkoxylation reaction. For either reaction, the catalyst tolerates a broad scope of substrates with various functional groups. By using the weakly electrophilic aluminum alkoxide as the catalyst, occurrence of “hidden Brønsted acid” catalysis can be excluded under the present reaction conditions.

Co-reporter:Dr. Lukas Hintermann;Dr. Jens Ackerstaff ;Dr. Florian Boeck
Chemistry - A European Journal 2013 Volume 19( Issue 7) pp:2311-2321
Publication Date(Web):
DOI:10.1002/chem.201203505

Abstract

Cinchona alkaloids catalyze the oxa-Michael cyclization of 4-(2-hydroxyphenyl)-2-butenoates to benzo-2,3-dihydrofuran-2-yl acetates and related substrates in up to 99 % yield and 91 % ee (ee=enantiomeric excess). Catalyst and substrate variation studies reveal an important role of the alkaloid hydroxy group in the reaction mechanism, but not in the sense of a hydrogen-bonding activation of the carbonyl group of the substrate as assumed by the Hiemstra–Wynberg mechanism of bifunctional catalysis. Deuterium labeling at C-2 of the substrate shows that addition of ROH to the alkenoate occurs with syn diastereoselectivity of ≥99:1, suggesting a mechanism-based specificity. A concerted hydrogen-bond network mechanism is proposed, in which the alkaloid hydroxy group acts as a general acid in the protonation of the α-carbanionic center of the product enolate. The importance of concerted hydrogen-bond network mechanisms in organocatalytic reactions is discussed. The relative stereochemistry of protonation is proposed as analytical tool for detecting concerted addition mechanisms, as opposed to ionic 1,4-additions.

Co-reporter:Lukas Hintermann;Claudia Dittmer
European Journal of Organic Chemistry 2012 Volume 2012( Issue 28) pp:5573-5584
Publication Date(Web):
DOI:10.1002/ejoc.201200838

Abstract

The asymmetric catalytic cyclization of the simple 2′-hydroxychalcone (1) to flavanone (2), a model for the chalcone isomerase reaction, has been realized as a catalytic asymmetric ion-pairing process with chiral quaternary ammonium salts (e.g., 9-anthracenylmethlycinchoninium chloride; 9-Am-CN-Cl) and NaH as small-molecule co-catalyst. In toluene/CHCl3 solution, the process reaches an intrinsic enantioselectivity of up to S = 14.4 (er = 93.5:6.5). The reversible reaction proceeds in two steps: A fast initial reaction approaches a quasi-equilibrium with KR/S = 4.5, followed by a second, slow racemization phase approaching Krac = 9. A simple mechanistic model featuring a living ion-pairing catalysis with full reversibility is proposed. Deuterium transfer from co-solvent CDCl3 to product 2 and isolation of a Michael conjugate formed from 2 and 1 demonstrate the intermediacy of flavanone enolate ion pairs. A kinetic model shows good agreement with the experimentally observed, peculiar, time-dependent evolution of the species concentrations and the enantiomeric excess of 2. The reaction is a chemical model of the chalcone isomerase enzymatic reaction. Furthermore, it is an ideal model for studying the characteristic behavior of reversible asymmetric catalyses close to their equilibria.

Co-reporter:Lukas Hintermann and Aleksej Turočkin
The Journal of Organic Chemistry 2012 Volume 77(Issue 24) pp:11345-11348
Publication Date(Web):November 20, 2012
DOI:10.1021/jo3021709
Addition of thioacetic acid to reactive α,β-unsaturated carbonyl compounds like acrolein or crotonaldehyde in acetone-d6 generates metastable (E)- and (Z)-1-alkenols, which tautomerize slowly at ambient temperature. The 1,4-addition of thioacetic acid and crotonaldehyde to (Z)-3-(acetylsulfanyl)-1-propen-1-ol is reversible with Keq = 5.5 ± 0.5 L/mol. A concerted, cyclic 1,4-addition mode is proposed to explain the preferred (Z)-stereoselectivity in lower polarity, nonprotic solvents.
Co-reporter:Dr. Lukas Hintermann
ChemCatChem 2012 Volume 4( Issue 3) pp:321-322
Publication Date(Web):
DOI:10.1002/cctc.201100407
Co-reporter:Florian Boeck ; Thomas Kribber ; Li Xiao
Journal of the American Chemical Society 2011 Volume 133(Issue 21) pp:8138-8141
Publication Date(Web):May 9, 2011
DOI:10.1021/ja2026823
The catalytic activity of [CpRu(L)2(MeCN)]PF6 (L = 2-diphenylphosphinopyridine with bulky groups at C-6) for anti-Markovnikov hydration of terminal alkynes to aldehydes is retained when one heterocyclic ligand L is replaced by L′ = PPh3. Equal amounts of CpRuCl(PPh3)2 (1) and phosphane L in acetone solution equilibrate to a mixture of 1, CpRuCl(L)(PPh3) (2), and CpRuCl(L)2 (3), which acts as highly active in situ catalyst for preparative anti-Markovnikov hydration of alkynes in water-rich media (2 mol % [Ru], 60 °C, 3–18 h in 4:1 (v/v) acetone/water). Reactions were completed in <15 min at 160 °C.
Co-reporter:Tuan Thanh Dang, Florian Boeck, and Lukas Hintermann
The Journal of Organic Chemistry 2011 Volume 76(Issue 22) pp:9353-9361
Publication Date(Web):October 19, 2011
DOI:10.1021/jo201631x
The generation of a hidden Brønsted acid as a true catalytic species in hydroalkoxylation reactions from metal precatalysts has been clarified in case studies. The mechanism of triflic acid (CF3SO3H or HOTf) generation starting either from AgOTf in 1,2-dichloroethane (DCE) or from a Cp*RuCl2/AgOTf/phosphane combination in toluene has been elucidated. The deliberate and controlled generation of HOTf from AgOTf and cocatalytic amounts of tert-butyl chloride in the cold or from AgOTf in DCE at elevated temperatures results in a hidden Brønsted acid catalyst useful for mechanistic control experiments or for synthetic applications.
Co-reporter:Lukas Hintermann;Marco Schmitz;Yun Chen
Advanced Synthesis & Catalysis 2010 Volume 352( Issue 14-15) pp:2411-2415
Publication Date(Web):
DOI:10.1002/adsc.201000350

Abstract

The nickel-catalyzed coupling of thiophene with aryl Grignard reagents (Wenkert reaction) is accelerated by N-heterocyclic carbene or trialkylphosphane ligands, providing a general, scalable direct synthesis of symmetrical 1,4-diarylbutadienes.

Co-reporter:Mateusz Rebarz, Boris-Marko Kukovec, Oleg V. Maltsev, Cyril Ruckebusch, Lukas Hintermann, Panče Naumov and Michel Sliwa
Chemical Science (2010-Present) 2013 - vol. 4(Issue 10) pp:NaN3809-3809
Publication Date(Web):2013/05/28
DOI:10.1039/C3SC50715G
The mysterious flashes of light communicated by fireflies conceal a rich and exciting solution spectrochemistry that revolves around the chemiexcitation and photodecay of the fluorophore, oxyluciferin. A triple chemical equilibrium by double deprotonation and keto–enol tautomerism turns this simple molecule into an intricate case where the relative spectral contributions of six chemical species combine over a physiologically relevant pH range, rendering physical isolation and spectral characterization of most of the species unmanageable. To disentangle the individual spectral contributors, here we demonstrate the advantage of chemical oriented multivariate data analysis. We designed a set of specific oxyluciferin derivatives and applied a multivariate curve resolution-alternating least squares (MCR-ALS) procedure simultaneously to an extensive set of pH-dependent spectroscopic data for oxyluciferin and the target derivatives. The analysis provided, for the first time, the spectra of the pure individual components free of contributions from the other forms, their pH-dependent profiles and distributions, and the most accurate to date values for the three equilibrium constants.
5-Pyrimidinecarboxaldehyde,2-(3-methoxyphenyl)-
2-(4-(trifluoromethyl)phenyl)pyrimidine-5-carbaldehyde
Benzene, [1-[[(1,1-dimethylethyl)dimethylsilyl]oxy]-3-butyn-1-yl]-
2-pyridin-4-ylpyrimidine-5-carbaldehyde
2-(4-fluorophenyl)pyrimidine-5-carbaldehyde
2-(4-Chlorophenyl)pyrimidine-5-carbaldehyde
2-(3-chlorophenyl)pyrimidine-5-carbaldehyde
2-(Thiophen-2-yl)pyrimidine-5-carbaldehyde