scispace - formally typeset
Search or ask a question

Showing papers in "Helvetica Chimica Acta in 1998"


Journal ArticleDOI
TL;DR: CD measurements, indicating a new secondary structure of certain beta-peptides constructed of beta2- and beta3-amino acids, were confirmed by detailed NMR soln.-structure anal, and possible factors influencing the relative stability of the two types of helixes are discussed.
Abstract: Enantiomerically pure beta-amino acid derivs. with the side-chains of Ala, Val, and Leu in the 2- or 3-position (β2- and β3-amino acids, resp.), as well as with substituents in both the 2- and 3-positions (β2,3-amino acids, of like-configuration) were prepd. and incorporated into β-hexa-, β-hepta-, and β-dodecapeptides. The new and some of the previously prepd. beta-peptides showed NH/ND exchange rates (in MeOH at room temp.) with tau1/2 ? 60 days, unrivaled by short-chain alpha-peptides. All beta-peptides were designed to be able to attain the previously described 31-helical structure. CD measurements, indicating a new secondary structure of certain beta-peptides constructed of beta2- and beta3-amino acids, were confirmed by detailed NMR soln.-structure anal. A beta2-heptapeptide and a beta2,3-hexapeptide have the 31-helical structure, while to a beta2/beta3-hexapeptide with alternating substitution pattern H-(beta2-Xaa-beta3-Xaa)3-OH a novel, unusual helical structure (in (D5)pyridine and in CD3OH) was assigned, with a central 10-membered and 2 terminal 12-membered H-bonded rings, and with C:O and N-H bonds pointing alternatively up and down along the axis of the helix. Thus, two types of beta-peptide turns were identified in soln. Hydrophobic interactions of and hindrance to solvent accessibility by the aliph. side-chains are discussed as possible factors influencing the relative stability of the two types of helixes.

269 citations


Journal ArticleDOI
TL;DR: The gamma-hexapeptide I (I) as discussed by the authors adopts a right-handed helical structure with a 3.61 helix of 5.4-ANG. pitch.
Abstract: For a comparison with the corresponding alpha- and beta-hexapeptides H-(Val-Ala-Leu)2-OH (A) and H-(beta-HVal-beta-HAla-beta-HLeu)2-OH (B), the corresponding gamma-hexapeptide (I), built from the homochirally similar (S)-4-aminobutanoate, (R)-4-amino-5-methylhexanoate, and (R)-4-amino-6-methylheptanoate, was prepd. The precursors were obtained either by double Arndt-Eistert homologation of protected Boc-Val-OH, Boc-Ala-OH, and Boc-Leu-OH, or by the superior route involving olefination/hydrogenation of the corresponding aldehydes, Boc-valinal, Boc-alaninal, and Boc-leucinal. Conventional peptide coupling (EDC/HOBt) furnished I through the intermediate gamma-di- and gamma-tripeptide. NMR measurements in (D5)pyridine and CD3OH soln. (COSY, TOCSY, HSQC, HMBC, ROESY) reveal that I adopts a right-handed helical structure [(P)-2.61 helix of .apprx.5-.ANG. pitch, contg. 14-membered H-bonded rings] which is to be compared with the left-handed helix of beta-peptide B [(M)-31 helix of 5-.ANG. pitch, 14-membered H-bonded rings] and with the familiar right-handed, so-called alpha-helix of alpha-peptides [(P)-3.61 helix of 5.4-.ANG. pitch, 13-membered rings]. Like the helix sense, the helix dipole reverses when going from alpha-(N->C) to beta-(C->N) to gamma-peptides (N->C). The surprising difference between the natural alpha-, and the analogous beta- and gamma-peptides is that the helix stability increases upon homologation of the residues.

199 citations


Journal ArticleDOI
TL;DR: The 4-isopropyl-5,5-diphenyloxazolidinone (1) is readily prepared from (R)- or (S)-valine ester, PhMgBr, and ethyl chlorocarbonate.
Abstract: The 4-isopropyl-5,5-diphenyloxazolidinone (1) is readily prepared from (R)- or (S)-valine ester, PhMgBr, and ethyl chlorocarbonate. It has a melting point of ca. 250°, a low solubility in most organic solvents, and a C=O group which is sterically protected from nucleophilic attack. Thus, the soluble N-acyl-oxazolidinones (7 – 16) can be prepared from 1 with BuLi at temperatures around 0° instead of −78° (Scheme 3), their Li enolates can be generated with BuLi, rather than with LDA, and deacylation in the final step of the procedure can be achieved with NaOH at ambient temperatures (Scheme 12), with facile recovery of the precipitating auxiliary 1 (filtering, washing, and drying). The following reactions of N-acyl-oxazolidinones from 1 have been investigated: alkylations (Scheme 4), aminomethylations and hydroxymethylations (Scheme 5), aldol additions (Schemes 6 and 7), Michael additions (Schemes 9 and 10), and a (4+2) cycloaddition (Scheme 11). The well-known features of reactions following the Evans methodology (yield, diastereoselectivity, dependence on conditions, counter ions, additives etc.) prevail in these transformations. Most products, however, have higher melting points and a much more pronounced crystallization tendency than those derived from conventional oxazolidinones, and can thus be purified by recrystallization, avoiding chromatography (Table 1). The disadvantage of 1 having a higher molecular weight (ca. 150 Da) than the non-phenyl-substituted auxiliary is more than compensated by the ease of its application, especially on large scale. A number of crystal structures of oxazolidinones derived from 1 and a TiCl4 complex of an oxazolidinone are described and discussed in view of the diastereoselective-reaction mechanisms.

178 citations


Journal ArticleDOI
TL;DR: The CD spectra of the β2- and β3-peptides 5, 9, and 10 show the typical pattern previously assigned to an (M) 31 helical secondary structure, which corresponds to a 2.5 fold increase in the molar ellipticity per residue!
Abstract: N-Fmoc-Protected (Fmoc = (9H-fluoren-9-ylmethoxy)carbonyl) β-amino acids are required for an efficient synthesis of β-oligopeptides on solid support. Enantiomerically pure Fmoc-β3-amino acids β3: side chain and NH2 at C(3)(= C(β)) were prepared from Fmoc-protected (S)- and (R)-α-amino acids with aliphatic, aromatic, and functionalized side chains, using the standard or an optimized Arndt-Eistert reaction sequence. Fmoc-β2- Amino acids (β2 side chain at C(2), NH2 at C(3)(= C(β))) configuration bearing the side chain of Ala, Val, Leu, and Phe were synthesized via the Evans' chiral auxiliary methodology. The target β3-heptapeptides 5–8, a β3- pentadecapeptide 9 and a β2-heptapeptide 10 were synthesized on a manual solid-phase synthesis apparatus using conventional solid-phase peptide synthesis procedures (Scheme 3). In the case of β3-peptides, two methods were used to anchor the first β-amino acid: esterification of the ortho-chlorotrityl chloride resin with the first Fmoc-β-amino acid 2 (Method I, Scheme 2) or acylation of the 4-(benzyloxy)benzyl alcohol resin (Wang resin) with the ketene intermediates from the Wolff rearrangement of amino-acid-derived diazo ketone 1 (Method II, Scheme 2). The former technique provided better results, as exemplified by the synthesis of the heptapeptides 5 and 6 (Table 2). The intermediate from the Wolff rearrangement of diazo ketones 1 was also used for sequential peptide-bond formation on solid support (synthesis of the tetrapeptides 11 and 12). The CD spectra of the β2- and β3-peptides 5, 9, and 10 show the typical pattern previously assigned to an (M) 31 helical secondary structure (Fig.). The most intense CD absorption was observed with the pentadecapeptide 9 (strong broad negative Cotton effect at ca. 213 nm); compared to the analogous heptapeptide 5, this corresponds to a 2.5 fold increase in the molar ellipticity per residue!

140 citations


Journal ArticleDOI
TL;DR: In this paper, a cyclophane-type molecular dyads was synthesized by Bingel macrocyclization of porphyrin-tethered bis-malonates 5 or 5.
Abstract: The synthesis of the cyclophane-type molecular dyads 1 and 1 . Zn was accomplished by Bingel macrocyclization of porphyrin-tethered bis-malonates 5 or 5 . Zn, respectively, with C60 (Scheme). In these macrocycles, the doubly bridged porphyrin adopts a close, tangential orientation relative to the surface of the C-sphere. The porphyrin derivatives 6 and 6 . Zn with two appended, singly-linked C60 moieties were also formed as side products in the Bingel macrocyclizations. The trans-1 addition pattern of the fullerene moiety in 1 and 1 . Zn was unambiguously established by 1H- and 13C-NMR spectroscopy. Due to the close spatial relationship between the fullerene and porphyrin components in 1 and 6 and the corresponding ZnII complexes, the porphyrin fluorescence is efficiently quenched as compared to the luminescence emitted by 5 and 5 . Zn, respectively (Fig. 2). Cyclic-voltammetry studies show that the mutual electronic effects exerted by the fullerene on the porphyrin and vice versa in 1 and 1 . Zn are relatively small despite the close proximity between the porphyrin donor and the fullerene acceptor (Fig. 3).

104 citations


Journal ArticleDOI
TL;DR: In this paper, the synthesis of β-peptides was reported, consisting of up to twelve β2, 2,2- or β3,3-geminally disubstituted β-amino acids which do not fit into any of the secondary structural patterns of β peptides, hitherto disclosed.
Abstract: We report on the synthesis of new and previously described β-peptides (1 – 6), consisting of up to twelve β2,2- or β3,3-geminally disubstituted β-amino acids which do not fit into any of the secondary structural patterns of β-peptides, hitherto disclosed. The required 2,2- and 3,3-dimethyl derivatives of 3-aminopropanoic acid are readily obtained from 3-methylbut-2-enoic acid and ammonia (Scheme 1) and from Boc-protected methyl 3-aminopropanoate by enolate methylation (Scheme 2). Protected (Boc for solution-, Fmoc for solid-phase syntheses) 1-(aminomethyl)cycloalkanecarboxylic-acid derivatives (with cyclopropane, cyclobutane, cyclopentane, and cyclohexane rings) are obtained from 1-cyanocycloalkanecarboxylates and the corresponding dihaloalkanes (Scheme 3). Fully 13C- and 15N-labeled 3-amino-2,2-dimethylpropanoic-acid derivatives were prepared from the corresponding labeled precursors (see asterixed formula numbers and Scheme 4). Coupling of these amino acids was achieved by methods which we had previously employed for other β-peptide syntheses (intermediates 18 – 23). Crystal structures of Boc-protected geminally disubstituted amino acids (16a – d) and of the corresponding tripeptide (23a), as well as NMR and IR spectra of an isotopically labeled β-hexapeptide (2a*) are presented (Figs. 1 – 4) and discussed. The tripeptide structure contains a ten-membered H-bonded ring which is proposed to be a turn-forming motif for β-peptides (Fig. 2).

96 citations


Journal ArticleDOI
TL;DR: For further structural studies and for physiological investigations of β-peptides, it is necessary to have H2O-soluble derivatives as mentioned in this paper, which can be used to detect possible π-π interactions.
Abstract: For further structural studies and for physiological investigations of β-peptides, it is necessary to have H2O-soluble derivatives Thus, we have prepared β-hexa-, β-hepta-, and β-nonapeptides (1 – 6) with two, three, and seven side chains of lysine and serine To detect possible π-π interactions, we also included the β-amino acid β2-HHop, resulting from homologation of so-called homophenylalanine (Hop) (5 and 6) The Fmoc-β2- and β3-amino-acid derivatives (11 – 14 and 19), and the corresponding β-peptides were prepared by methods previously described (solid-phase peptide coupling; HPLC-pure samples, Fig 1) Circular-dichroism spectra (Fig 2) indicate the presence of less pronounced secondary structures (especially of the lysine analogues with multiple positive charge) in H2O as compared to MeOH The β3-heptapeptide (3) with two serine side chains is well soluble in H2O and exhibits the CD pattern typical of the 31-helical structure

81 citations


Journal ArticleDOI
TL;DR: A comprehensive survey of the relationship between purine-purine pairing in homo-DNA oligonucleotides is given in this article, together with a survey of what we know today about pairing in all possible purinepurine combinations.
Abstract: Why Pentose- and Not Hexose-Nucleic Acids? Purine-Purine Pairing in homo-DNA: Guanine,Isoguanine, 2,6-Diaminopurine, and Xanthine This paper concludes the series of reports in this journal [1–4] on the chemistry of homo-DNA, the constitutionally simplifie dmodel system of hexopyranosyl-(6′ → 4′)-oligonucleotide systems stidued in our laboratory as potentially natural-nucleic-acid alternatives in the context of a chemical aetiology of nucleic-acid structure. The report describes the synthesis and pairing properties of homo-DNA oligonucleotides which contain as nucleobases exclusively purines, and gives, together with part III of the series [3], a survey of what we know today about purine-purine pairingin homo-DNA. In addition, the paper discusses those aspects of the chemistry of homo-DNA which, we think, influence the way how some of the structural features of DNA (and RNA) are to be interpreted on a qualitative level. Purine-purine pairing occurs in the homo-DNA domain in great variety. Most prominent is a novel tridentate Watson-Crick pair between guanine and isoguanine, as well as one between 2,6-diaminopurine and xanthinone, both giving rise to very stable duplexes containing the all-purine strands in antiparallel orientation. For the guanine-isoguanine pair, constitutional assignment is based on temperature-dependent UV and CD spectroscopy of various guanine- and isoguanine-containg duplexes in comparison with duplexes known to be paired in the reverse guanine is replaced by 7-carbauguanine. Isoguanine and 2,6-diaminopurine also have the capability of self-pariring in the reverse-Hoogsteen mode, as previously observed for adenine and guanine [3]. In this type of pairing, the interchangeably. Fig. 36 provides an overall survey of the relative strength of pairing in all possible purine-purine combinations. Watson-Crick pairing of isoguanine with guanine demands the former to participate in its 3H-tautomeric form; hitherto this specific tautomer had not been considered in the pairing chemistry of isoguanine. Whereas (cumulative) purine-purine pairing in DNA (reverse-Hoogsten or Hoogsteen) seems to occur in triplexes and tetrapalexes only, its occurrence in duplexes in a characteristic feature of homo-DNA chemistry. The occurrence of purine-purine Watson-Crick base pairs is probably a consequence of homo-DNA's quasi-linear ladder structure [1][4]. In a double helix, the distance between the two sugar C-atoms, on which a base pair is anchored, is expected to be constrained by the dimensions of the helix; in a linear duplex, however, there would be no restrictions with regard to base-pair length. Homo-DNA's ladder-like model also allows one to recognize one of the reasons why nucleic-acid duplexes prefer to pair in antiparallel, rather than parallel strand orientation: in homo-DNA duplexes, (averaged) backbone and base pair axes are strongly inclined toward one another [4]; the stronger this inclination, the higher the preference for antiparallel strand orientation is expected to be (Fig. 16). In retrospect, homo-DNA turns out to be one of the first artificial oligonucleotide systems (cf. Footnote 65) to demonstrate in a comprehensive way that informational base pairing involving purines and pyrimidines is not a capability unique to ribofuranosyl systems. Stability and helical shape of pairing complexes are not necessary conditions of one another; it is the potential for extensive conformational cooperativity of hte backbone structure with respect to the constellational demands of base pairing and base stacking that determines whether or nor a given type of base-carrying backbone structure is an informational pairing system. From the viewpoint of the chemical aetiology of nucleic-acid structure, which inspired our investigations on hexopyranosyl-(6′ → 4′)-oligonucleotide systems in the first place, the work on homo-DNA is only an extensive model study, because homo-DNA is not to be considered a potential natural-nucleic-acid altenratie. In retrospect, it seems fortunate that the model study was carried out, because without it we could hardly have comprehended the pairing behavior of the proper nucleic-acid alternatives which we have studied later and which will be discussed in Part VI of this series. The English footnotes to Fig. 1–49 provide an extension of this summary.

78 citations


Journal ArticleDOI
TL;DR: In this paper, X-ray analysis of the platelets obtained from CHCl3 reveals a disordered fibre-like structure consisting of stacks of ribbon-like aggregates, which are completely different from the columnar structures, based on G-quartets, which were the building blocks of the mesophases formed by deoxyguanosine oligonucleotides in H2O.
Abstract: The lipophilic 3′,5′-di-O-decanoyl-2′-deoxyguanosine (1) in CHCl3 undergoes extensive self-assembly, mediated by H-bonding between the guanine bases, to give ribbon-like aggregates. X-Ray investigation of the platelets obtained from CHCl3 reveals a disordered fibre-like structure consisting of stacks of the ribbon-like aggregates. The aggregates are completely different from the columnar structures, based on G-quartets, which are the building blocks of the mesophases formed by deoxyguanosine oligonucleotides in H2O. In pure hydrocarbons or in CHCl3/hydrocarbons, 1 forms a lyotropic liquid-crystalline phase.

66 citations


Journal ArticleDOI
TL;DR: In this paper, a new family of optically active cyclophane receptors, (1, 1, 3)-diynediyl linkers, is described, which contain highly preorganized cavities lined with six convergent OH groups for H-bonding and complementary in size and shape to monosaccharides.
Abstract: The synthesis and carbohydrate-recognition properties of a new family of optically active cyclophane receptors, 1 – 3, in which three 1,1′-binaphthalene-2,2′-diol spacers are interconnected by three buta-1,3-diynediyl linkers, are described. The macrocycles all contain highly preorganized cavities lined with six convergent OH groups for H-bonding and complementary in size and shape to monosaccharides. Compounds 1 – 3 differ by the functionality attached to the major groove of the 1,1′-binaphthalene-2,2′-diol spacers. The major grooves of the spacers in 2 are unsubstituted, whereas those in 1 bear benzyloxy (BnO) groups in the 7,7′-positions and those in 3 2-phenylethyl groups in the 6,6′-positions. The preparation of the more planar, D3-symmetrical receptors (R,R,R)-1 (Schemes 1 and 2), (S,S,S)-1 (Scheme 4), (S,S,S)-2 (Scheme 5), and (S,S,S)-3 (Scheme 8) involved as key step the Glaser-Hay cyclotrimerization of the corresponding OH-protected 3,3′-diethynyl-1,1′-binaphthalene-2,2′-diol precursors, which yielded tetrameric and pentameric macrocycles in addition to the desired trimeric compounds. The synthesis of the less planar, C2-symmetrical receptors (R,R,S)-2 (Scheme 6) and (S,S,R)-3 (Scheme 9) proceeded via two Glaser-Hay coupling steps. First, two monomeric precursors of identical configuration were oxidatively coupled to give a dimeric intermediate which was then subjected to macrocyclization with a third monomeric 1,1′-binaphthalene precursor of opposite configuration. The 3,3′-dialkynylation of the OH-protected 1,1′-binaphthalene-2,2′-diol precursors for the macrocyclizations was either performed by Stille (Scheme 1) or by Sonogashira (Schemes 4, 5, and 8) cross-coupling reactions. The flat D3-symmetrical receptors (R,R,R)-1 and (S,S,S)-1 formed 1 : 1 cavity inclusion complexes with octyl 1-O-pyranosides in CDCl3 (300 K) with moderate stability (ΔG0ca. −3 kcal mol−1) as well as moderate diastereo- (Δ(ΔG0) up to 0.7 kcal mol−1) and enantioselectivity (Δ(ΔG0)=0.4 kcal mol−1) (Table 1). Stoichiometric 1 : 1 complexation by (S,S,S)-2 and (S,S,S)-3 could not be investigated by 1H-NMR binding titrations, due to very strong signal broadening. This broadening of the 1H-NMR resonances is presumably indicative of higher-order associations, in which the planar macrocycles sandwich the carbohydrate guests. The less planar C2-symmetrical receptor (S,S,R)-3 formed stable 1 : 1 complexes with binding free enthalpies of up to ΔG0=−5.0 kcal mol−1 (Table 2). With diastereoselectivities up to Δ(ΔG0)=1.3 kcal mol−1 and enantioselectivities of Δ(ΔG0)=0.9 kcal mol−1, (S,S,R)-3 is among the most selective artificial carbohydrate receptors known.

63 citations


Journal ArticleDOI
TL;DR: In this article, a cycloadduct of type 7 via zwitterion was proposed, and a reaction mechanism was proposed by a two-step formation of the formal cycle-oadduct.
Abstract: The reactions of 1,4,5-trisubstituted imidazole 3-oxides 1a - k with cyclobutanethiones 5a,b in CHCl3 at room temperature give imidazole-2(3H)-thiones 9a - k in high yield. The second product formed in this reaction is 2,2,4,4-tetramethylcyclobutane-1,3-dione (6a; Scheme 2). Similar reactions occur with 1 and adamantanethione (5c) as thiocarbonyl compound, as well as with 1,2,4-triazole-4-oxide derivative 10 and 5a (Scheme 3). A reaction mechanism by a two-step formation of the formal cycloadduct of type 7 via zwitterion 16 is proposed in Scheme 5. Spontaneous decomposition of 7 yields the products of this novel sulfur-transfer reaction. The starting imidazole 3-oxides are conveniently prepared by heating a mixture of 1,3,5-trisubstituted hexahydro-1,3,5-triazines 3 and alpha-(hydroxyimino) ketones 2 in EtOH (cf. Scheme 1). As demonstrated in the case of 9d, a 'one-pot' procedure allows the preparation of 9 without isolation of the imidazole 3-oxides 1. The reaction of 1c with thioketene 12 leads to a mixture of four products (Scheme 4). The minor products, 9c and the ketene 15, result from an analogous sulfur-transfer reaction (Path a in Scheme 5), whereas the parent imidazole 14 and thiiranone 13 are the products of an oxygen-transfer reaction (Path b in Scheme 5).

Journal ArticleDOI
TL;DR: The α- and γ-CD analogues were synthesized by intramolecular coupling of the bis-O-propargylated maltohexaoside 4, or the analogous maltooctaoside 16, followed by deprotection as mentioned in this paper.
Abstract: The α- and γ-CD analogues 6 and 18, which possess a hexa-2,5-diyne-1,6-dioxy unit, were synthesised by intramolecular coupling of the bis-O-propargylated maltohexaoside 4, or the analogous maltooctaoside 16, followed by deprotection. The dialkynylated linear oligosaccharides were obtained by glycosidation of propargyl alcohol with the thioglycosides 1 and 13, reductive cleavage of the benzylidene acetal, and propargylation of the terminal HO−C(4) group, respectively. The β-CD analogues 23 and 25, which possess a penta-1,3-diyn-1-yl-5-oxy unit, were similarly obtained by intramolecular oxidative coupling of 20 and 21, respectively. The linear dialkynylated oligosaccharides 20 and 21 were obtained by two consecutive glycosylations, first with the maltohexaosyl-S-glycoside 1 as donor, and then by glycosylation of the resulting propargyl maltohexoside with the C(4)-ethynylated donor 19. The proximity of the terminal units of maltooligosaccharides allowed a facile intramolecular cycloaddition of the azido alkyne 29 to the isomeric triazoles 30 and 31, which were deprotected to 32 and 33, respectively. Analysis of the intramolecular H-bonds in 6, 23, 25, 32, and 33 showed that insertion of a noncarbohydrate link interrupts a single flip-flop H-bond.

Journal ArticleDOI
TL;DR: In this paper, the crystal structures of the supermolecular pairs I-III and of the homochiral aggregate of two ternary super-molecules IV have been determined.
Abstract: Chiral supermolecules may be obtained from suitable achiral molecular constituents associated through a dissymmetrizing interaction mode. This is the case for the supermolecules I–IV formed by hydrogen-bonding association between the achiral complementary components 1a,b and 2a,b,c. The crystal structures of the supermolecular pairs I–III and of the homochiral aggregate of two ternary supermolecules IV have been determined. The structural data are discussed.

Journal ArticleDOI
TL;DR: The first N,N-dialkyl carbamato complex of gold was reported in this article, which consists of dinuclear [(Ag2OCNMe2)2] units with slightly distorted linearly two-coordinated Ag-atoms containing bridging caramato groups to form a substantially planar eight-membered ring.
Abstract: The reaction of CuO'Bu with CO2, and iPr2NH in the presence of PPh3, gives the dialkylcarbamato complex [Cu(O2CNiPr2)(PPh3)2] (1). The CO2/R2NH system (R = Me, Et) in an appropriate organic medium reacts with Ag2O giving the corresponding N,N-dialkylcarbamato complexes of analytical formula [Ag(C2CNR2)] (R = Me, 2; R = Et, 3). The methyl derivative 2 was characterized by X-ray diffraction methods. Crystal data of 2: for [Ag2(O2CNMe2)2], C6H12Ag2N2O4, mol. wt. 391.9; monoclinic, space group P21/c, a = 12.08(1), b = 3.797(2), c = 11.316(7) A, β = 113.37(6)°, V = 476.3 A3, Z = 2, Dc = 2.732 g cm−3; μ(MoKα) = 40.64 cm−1, F(000) = 376.0; R = 0.059, Rw = 0.067; g.o.f. 1.27. The structure consists of dinuclear [(Ag2OCNMe2)2] units with slightly distorted linearly two-coordinated Ag-atoms containing bridging carbamato groups to form a substantially planar eight-membered ring with an intra-annular AgAg distance of 2.837(2) A; the dinuclear units are further joined by AgO bonds to form an infinite array. Compound 3, which is presumably dinuclear, as suggested by cryoscopic measurements in benzene, undergoes a structural fission with PPh3, giving the mononuclear triphenylphosphine derivative [Ag(O2CNEt2)(PPh3)2] (4). The amine-catalyzed conversion of Ag2O into Ag2CO3, in the presence of the iPr2NH/CO2 system, is also reported. Cl-Exchange from [AuCl(PPh3)] with [Ag(O2CNEt2)] (3) gives the first N,N-dialkylcarbamato complex of gold, namely [Au(O2CNEt2)(PPh3)] (5), which crystallizes in the monoclinic system: C23H25AuNO2P · 0.5 C7H16, mol. wt. 625.5, space group P21/c; a = 13.212(5), b = 12.25(1), c = 16.795(6) A, β = 109.09(2)°, V = 2568(2) A3, Z = 4, Dc, = 1.618 g cm−3; μ(AgKα) = 31.40 cm−1, F(000) = 1236.0; R = 0.058; Rw = 0.064; g.o.f. 2.121. The compound contains two-coordinated Au-atom, namely to the P-atom and to the O-atom of the monodentate carbamato group, the PAuO bond angle being 174.7(3)°. The reaction with MeI showed these compounds to react predominantly at the carbamato O-atom giving the corresponding urethanes R2NCO2Me. Evidence was gathered for the transient coordination of CO to Ag in 3.

Journal ArticleDOI
TL;DR: The templating effect of the heterochiral diproline unit also promotes efficient backbone cyclization of the fourteen-residue linear peptide precursor, suggesting that a wide variety of related protein loop mimetics incorporating the D-Pro-L-Pro template might be readily accessible.
Abstract: A straightforward and effective method of stabilizing a β-hairpin conformation in a cyclic protein loop mimetic is described, which exploits the templating effect of a heterochiral D-Pro-L-Pro dipeptide unit. A twelve-residue β-hairpin loop was grafted from the extracellular interferon γ receptor onto the heterochiral D-Pro-L-Pro dipeptide template to afford a fourteen-residue cyclic peptide. The residues directly attached to the D-Pro-L-Pro template are shown by NMR spectroscopy to structurally mimic corresponding residues in adjacent antiparallel β-strands in the receptor. MD Simulations with and without time-averaged distance restraints support this view and indicate that the tip of the loop is more flexible, as inferred also for the receptor protein from crystallographic data. The templating effect of the heterochiral diproline unit also promotes efficient backbone cyclization of the fourteen-residue linear peptide precursor, suggesting that a wide variety of related protein loop mimetics incorporating the D-Pro-L-Pro template might be readily accessible.

Journal ArticleDOI
TL;DR: In this article, a series of chiral enantiomerically pure C2-symmetric binucleating ligands were synthesized from chiral amino alcohols, which can coordinate two metals in close proximity to each other, are of interest regarding possible applications in asymmetric catalysis.
Abstract: The synthesis of a series of chiral enantiomerically pure C2-symmetric binucleating ligands is reported. Ligands of type 1–4, which consist of a phenolic or heterocyclic unit bridging two chiral dihydrooxazole rings. are readily accessible from chiral amino alcohols. Ligands 5a and 5b are composed of a cyclic urea or thiourea unit, respectively, and two 3,4-dihydro-2H-pyrrole rings containing a stereogenic center next to the N-atom. Compounds of this type are readily assembled from ethane-1,2-diamine and an imidothioic ester derived from pyroglutamic acid. These new ligands, which can coordinate two metals in close proximity to each other, are of interest regarding possible applications in asymmetric catalysis.

Journal ArticleDOI
TL;DR: In this article, two series of linearly conjugated hybrid materials, consisting of (E)-1,2-diethynylethene (DEE; hex-3-ene-1,5-diyne) and ZnII porphyrin components, were prepared by Pd0-catalyzed cross-coupling reactions.
Abstract: Two series of linearly conjugated hybrid materials, consisting of (E)-1,2-diethynylethene (DEE; hex-3-ene-1,5-diyne) and ZnII porphyrin components, were prepared by Pd0-catalyzed cross-coupling reactions. In one series, one or two DEE substituents were introduced into the meso-positions of the ZnII porphyrins, leading from 5⋅Zn, to 9 and 1 (Scheme 1). The second series contains the linearly π-conjugated molecular rods 1 – 3 that span a length range from 23 A (1) to 53 A (3) (Fig. 1). The larger rods 2 and 3 consist of two or three porphyrin moieties, respectively, that are bridged at the meso-positions by trans-enediynediyl (hex-3-ene-1,5-diyne-1,6-diyl) linkers (Scheme 2). The UV/VIS spectra in the series 5⋅Zn, 9, and 1 (Fig. 2) showed a strong bathochromic shift of both Soret and Q bands of the ZnII porphyrin as a result of the addition of DEE substituents. Upon changing from 1 to 2 (Fig. 3), the Q band was further bathochromically shifted, whereas the Soret band remained nearly at the same position but became broadened and displayed a shoulder on the lower-wavelength edge as a result of excitonic coupling. The close resemblance between the UV/VIS spectra of 2 and 3 suggests that saturation of the optical properties in the oligomeric series already occurs at the stage of dimeric 2. Stationary voltammetric investigations showed that the DEE substituents act as strong electron acceptors which induce large anodic shifts in the first reduction potential upon changing from 5⋅Zn to 9 (ΔE=190 mV) and to 1 (ΔE=340 mV). Increasing the number of porphyrin moieties upon changing from 1 to 2 had no effect on the first reduction potential yet the first oxidation potential was substantially lowered (ΔE=110 mV). Large differences in the potentials for one-electron oxidation of the two porphyrin moieties in 2 (ΔE=200 mV) confirmed the existence of substantial electronic communication between the two macrocycles across the trans-enediynediyl bridge.

Journal ArticleDOI
TL;DR: In this article, the relative configuration of the peroxide/α-methylacetate moiety of 6, 9, and 10, was directly determined from their NMR spectra.
Abstract: Natural free carboxylic acids from the hadromerid sponge Diacarnus levii (Kelly-Borges and Vacelet) were esterified to yield the new cyclic norditerpene peroxides ent-muqubilin benzyl ester (= (αR,3S,6R)-α,6-dimethyl-6-[(E)-4-methyl-6-(2,6,6-trimethyl-cyclohex-1-en-1-yl)hex-3-enyl]-1,2-dioxan-3-acetic acid benzyl ester; 6, diacarnoate B methyl ester(= (αS,3R,6R)-α,6-dimethyl-6-{2-(4aS,8aS)-3,4,4a,5,6,7,8,8a-octahydro-3-oxo-2,5,5,8a-tetramethylnaphthalen-1-yl)ethyl}-1,2-dioxan-3-acetic acid methyl ester; 9), and deoxydiacarnoate B benzyl ester (= (αS,3R,6R)-α,6-dimethyl-6-{2-[(4aS,8aS)-3,4,4a,5,6,7,8,8a-octahydro-2,5,5,8a-tetramethyl-1-naphthalenyl]ethyl}-1,2-dioxan-3-acetic acid benzyl ester; 10), which were isolated following extensive chromatography. The relative configuration of the peroxide/α-methylacetate moiety of 6, 9, and 10, was directly determined from their NMR spectra. The absolute configurations of the peroxide/α-methylacetate moiety was deduced from comparative 1H-NMR data of the (S)- and (R)-phenylglycine methyl ester derivatives 7 and 8 as well as 11/13 and 12/14, all obtained from a mixture of the precursors of 3, 6, and 10. The absolute configuration at the carbobicyclic moiety of enone 9 and of 10, is identical, as established by chemical interconversion, 9 and 10 belong to the normal labdane series according to empirical CD rules, applied either directly to 9 or to the parent (+)-sclareolide-derived enone 20. In contrast, molar rotation additivity rules suggest the ent-labdane configuration for 9 and 10. The epidioxides 1–3, 6, and 10 proved active in vitro against the malaria parasite Plasmodium falciparum; especially the previously isolated methyl 3-epinuapapuanoate (2) was active against a chloroquine-resistant strain, and this with a good security index.

Journal ArticleDOI
TL;DR: In this paper, a novel general synthesis of 2,4,5-tri-and 2, 4,5,6-tetrasubstituted pyrimidines by condensation of thiouronium salts of type 3 with (ethoxymethylidene)malononitrile (4) and [bis(methylthio)methylidene]malononiitrile(6), respectively, was first established in solution and successfully transferred onto solid support by using the polymer-bound pyrimidine salt.
Abstract: A novel general synthesis of 2,4,5-tri- and 2,4,5,6-tetrasubstituted pyrimidines 5a–d and 7a, e, f, g by condensation of thiouronium salts of type 3 with (ethoxymethylidene)malononitrile (4) and [bis(methylthio)methylidene]malononitrile (6), respectively, was first established in solution (Scheme 1) and successfully transferred onto solid support by using the polymer-bound thiouronium salt 11 (Scheme 3) Further investigations were directed toward a multidirectional cleavage procedure of the 2-(alkylsulfinyl) intermediates, obtained from the 2-(alkylthio)pyrimidines 7a (Scheme 2) or 12 and 14 (Schemes 3 and 4), with different nucleophiles to form highly substituted pyrimidines In addition, fused-heterocycle derivatives 22a--h, 24a–c, and 26a–e were generated in good-to-excellent yields by condensation of 7a, e, h with versatile isocyanates and isothiocyanates, with subsequent alkylation (Scheme 5)


Journal ArticleDOI
TL;DR: In this article, a rational synthetic design at the ferrocene level led to columnar liquid-crystals with columnar behavior and to H-bonded metallomesogens.
Abstract: The synthesis and liquid-crystalline properties of tetracatenar covalent and H-bonded bis-ferrocene derivatives 1 and 2, respectively, are reported. Both compounds gave rise to enantiotropic columnar liquid- crystalline behavior with a hexagonal molecular organization. To explore the possibility to obtain also calamitic liquid-crystalline phases from H-bonded ferrocene-containing liquid crystals, a rod-shaped ferrocene mesogen 3 was synthesized, which gave rise to enantiotropic smectic C and smectic A phases. For the first time, a rational synthetic design at the ferrocene level led to ferrocene-based liquid-crystals with columnar behavior and to H- bonded metallomesogens. Introduction. - Current efforts oriented towards the development of ferrocene- based mesomorphic materials led to the synthesis of ferrocene-containing side-chain liquid-crystalline polymers (1), ferrocene-containing liquid-crystalline dendrimers (2), and a liquid-crystalline mixed (60)fullerene-ferrocene derivative (3). These structures are promising candidates to construct switchable anisotropic materials. We demon- strated that electron transfer can be used to generate mesomorphism in the ferrocene- ferrocenium redox system (4). To further explore the structure/supramolecular organization relationship in the case of ferrocene-containing mesomorphic molecular units, the study of novel structures is required. Mesomorphic ferrocenes exhibiting columnar liquid-crystalline phases would be of particular interest: No such examples have been reported (5 - 7) 1 ). Because of the bulky, three-dimensional structure of the ferrocene core, the primary influence of which is to reduce the intermolecular interactions (resulting in a lowering of the liquid-crystalline tendency in comparison with the ferrocene-free analogues) (5), the design of such compounds represents both a conceptual and synthetic challenge. Ultimately, owing to its redox activity, the incorporation of ferrocene units into columnar phases may open the doors to new supramolecular wires (9).

Journal ArticleDOI
TL;DR: The cyclodextrin ‘dimer’ 4 was obtained via condensation of the corresponding 4,4′-(porphyrin-5,15-diyl)bis phenol 8 with 6A-deoxy-6A-iodo-β-cyclodextrine in the presence of Cs2CO3.
Abstract: We report on the synthesis of a receptor 4 for β,β-carotene (1), and on the binding interaction between the two which yields inclusion complex 5. The cyclodextrin ‘dimer’ 4 was obtained via condensation of the corresponding 4,4′-(porphyrin-5,15-diyl)bis phenol 8 with 6A-deoxy-6A-iodo-β-cyclodextrin (6) in the presence of Cs2CO3. Fluorescence studies of the binding interaction between ‘dimer’ 4 and β,β-carotene (1) gave a binding constant Ka of (2.4 ± 0.1) · 106M−1.

Journal ArticleDOI
TL;DR: In this article, a broad variety of acyclic vinyl ethers have been synthesized via the vinyl-interchange reaction of ethyl vinyl ether at room temperature using mercury(II) trifluoroacetate as a highly efficient catalyst.
Abstract: A broad variety of new acyclic vinyl ethers (see 6–41) have been synthesized via the vinyl-interchange reaction of ethyl vinyl ether at room temperature using mercury(II) trifluoroacetate as a highly efficient catalyst. The appropriate vinyl ethers were reacted under acidic conditions with 3′,5′-O-silyl-protected uridine 42 to the corresponding 2′-O-(1-alkoxyethyl) derivatives 43–83 which gave, on desilylation of F− ions, in high yields the uridine-2′-O-acetal derivatives 84–124. The relative stabilities of the newly synthesized compounds under acidic and basic conditions were determined using TLC and HPLC techniques. Protected protecting groups offer the best properties for oligoribonucleotide syntheses. Interestingly, the very acid-stable acetals of the β-substituted ethyl-type 118–121 and 123 can be cleaved by a β-elimination process providing a series of base-labile acetals of potential synthetic value.

Journal ArticleDOI
TL;DR: In this paper, the Dess-Martin periodinane, acid-catalysed isomerization of the oxidation products in toluene, and reductive dehydroxylation of 18/19 (Et3SiH/BF3 · OEt2; Scheme 1).
Abstract: The lactam 21 was obtained in an overall yield of 72% from the hydroxy amide 16 by oxidation with the Dess-Martin periodinane, acid-catalysed isomerization of the oxidation products in toluene, whereupon 18/19 precipitated, and reductive dehydroxylation of 18/19 (Et3SiH/BF3 · OEt2; Scheme 1). The amide 16 was obtained by ammonolysis of the N-acetylglucosamine-derived lactone 15. Depending on the oxidation method, 16 yielded the keto amide 17, the hydroxy lactams 18/19, and the pyrrolidinecarboxamide 20 in widely different proportions. The pyrrolidinecarboxamide 20 was not reduced under the conditions of the reductive dehydroxylation. Hydrogenolysis of the benzyl-protected lactam 21 gave the trihydroxy lactam 22, while reduction with NaBH4/BF3 · OEt2 led to the 2-acetamidopiperidine derivative 24 (Scheme 2). Selective (tert-butoxy)carbonylation of the lactam 21 ( 25) followed by NaBH4 reduction and acid-catalysed solvolysis in EtOH led to the α-ethoxycarbamates 28/29. Similarly, (tert-butoxy) carbonylation of 1 ( 31) followed by reduction to 32/33 and glycosidation yielded the ethoxycarbamate 34. Treatment of the GlcNAc-derived ethyl glycosides 28/29 with Me3SiCN/BF3 · OEt2 gave the equatorial amino nitrile 30. Under similar conditions, the Glc-derived glycoside 34 led to the iminooxazolidinone 35. In the presence of a larger proportion of Me3SiCN at 5°, 34 was transformed into the axial, selectively monodebenzylated amino nitrile 36.

Journal ArticleDOI
TL;DR: In this paper, the antifungal activity of the CH2Cl2 extract of Monotes engleri Gilg. (Dipterocarpaceae) was investigated in bioautographic TLC assays.
Abstract: The CH2Cl2 extract of Monotes engleri Gilg. (Dipterocarpaceae) showed antifungal activity against the yeast Candida albicans in our bioautographic TLC assays. After a first fractionation of the crude extract, the bioactivity was located in one of the fractions. To perform an efficient targeted isolation of the active compounds, LC/UV/MS and LC/UV/NMR analyses of the crude extract and the active fraction were performed. LC/UV/, LC/MS, and LC/NMR data (1D and 2D) allowed the identification of 1 as (2S)-2,3-dihydro-5,7-dihydroxy–{3-hydroxy-4-[(3-methylbut-2-enyl) oxy]phenyl}-4H-1-benzopyran-4-one, a new prenylated flavanone, named monoteson A. Subsequent isolation of 1 has permitted the determination of its absolute configuration on the basis of CD measurements. Theree other prenylated flavanoes 2–4 were isolated from the same extract. Compound 3 was identified as 2- (3, 5-dihydroxyphenyl) -2,3-dihydro-5, 7-dihydroxy-6, 8-bis (3-methylbut-2-enyl)-4 H-1-benzopyran-4-one, another new natural product, named monotesone B. The structures of 2 and 4 were established as selinone and lonchocarpol A, respectively. The antifungal activity against Candida albicans was determined for all compounds.

Journal ArticleDOI
TL;DR: The first steps involved in the highly enantioselective copolymerization of styrene and carbon monoxide to poly[1-oxo-2-phenylpropane-1,3-diyl] with phosphinodihydrooxazole-palladium(II) complexes were investigated in this paper.
Abstract: The first steps, believed to be involved in the highly enantioselective copolymerization of styrene and carbon monoxide to poly[1-oxo-2-phenylpropane-1,3-diyl] with phosphinodihydrooxazole-palladium(II) complexes, were investigated. The insertion of carbon monoxide into [Pd(Me)(PˆN)(solvent)] TfO (PˆN = (S)-2-[2-(5H-benzo[b]phosphindol-5-yl)phenyl]4-benzyl-4,5-dihydrooxazole (1)) and of styrene into [Pd(Me)(PˆN)(solvent)] TfO were highly regioselective (alkyl and acyl substituents trans to N); moreover, the olefin insertion took place with essentially complete enantioface discrimination.

Journal ArticleDOI
TL;DR: In spite of the higher reactivity of styrene with respect to ethene for the alternating copolymerization with carbon monoxide, catalyzed by chiral (dihydrooxazole)(phosphino)palladium complexes, Ethene is preferentially (and randomly) enchained in terpolymerization experiments; enantioface selection for styrene is comparably high in both copolymersization and ter polymerization processes as mentioned in this paper.
Abstract: In spite of the higher reactivity of styrene with respect to ethene for the alternating copolymerization with carbon monoxide, catalyzed by chiral (dihydrooxazole)(phosphino)palladium complexes, ethene is preferentially (and randomly) enchained in terpolymerization experiments; enantioface selection for styrene is comparably high in both copolymerization and terpolymerization processes.

Journal ArticleDOI
TL;DR: In this paper, the synthesis of 3H-naphtho[2,1-b]pyrans 9−21 linked to a thiophene moiety is described.
Abstract: The synthesis of 3H-naphtho[2,1-b]pyrans 9–21 linked to a thiophene moiety is described. Two different synthetic approaches were applied to prepare these novel functionalized compounds, and their spectrokinetic properties in solution are reported.

Journal ArticleDOI
TL;DR: In this article, a variety of 7-O-substituted analogs of deaminodeoxycolchinol thiomethyl ether were synthesized and evaluated for their inhibitory effects on tubulin polymerization in vitro.
Abstract: 7-O-Substituted analogs of deaminodeoxycolchinol thiomethyl ether were synthesized and evaluated for their inhibitory effects on tubulin polymerization in vitro. Ketone 9, a key compound in this study, was derived from thiocolchicone 6 by reaction with aniline. Reaction of compound 6 with MeNH2 or BuNH2 gave tetracyclic lactams 7 and 8, respectively. Optically active alcohols 11a and 11b were obtained from racemic 11 by chemical resolution including a separation of the comphanate diastereoisomers 12a and 12b, followed by basic hydrolysis. The (aR,7R)-configuration of 12b was verified by X-ray crystallographic analysis. Almost all racemic and optically active 7-O-acyl or 7-O-aroyl compounds had strong inhibitory effects on the tubulin polymerization reaction, with IC50 values from 1.7 to 5.1 μM. A few agents, such as the lactams 7 and 8, the camphanates 12a and 12b, the cyclohexanecarboxylates 19a and 19b, and, most notably, the (7S)-benzoate 15a, had negligible effects on polymerization, yielding IC50 values greater than 40μM. Ketone 9 showed strong inhibition of tubulin polymerization comparable to that of thiocolchicone (6). Optically active alcohol 11a and acyl esters 13a and 14a with a (7S)-configuration were more active than the (7R)-esters 13b and 14b. However, the esters 15a–17a with a (7S)-configuration were less active than the (7R)-isomers 15b–17b, in which the (7R)-benzoate 15b was at least 15-fold more inhibitory than the (7S)-isomer 15a. For the most part, the agents causing strongest inhibition of polymerization also caused the greatest inhibition of [3H]colchicine binding. NMR and optical rotatory data indicate that, in polar solvents, the equilibrium in esters with a 7-O-aroyl substituent, i.e., 15a,b, 16a,b, and 17a,b, is reversed from (aS) to (aR) or from (aR) to (aS), as compared to nonpolar solvents.

Journal ArticleDOI
TL;DR: New derivatives of campholenaldehyde (2,2,3-trimethylcyclopent-3-enyl) bearing two cyclopropane moieties were synthesized, and the structure of the stereoisomer responsible for its exceptionally strong, diffusive, and natural sandalwood-oil scent was elucidated.
Abstract: New derivatives of campholenaldehyde (= 2-(2,2,3-trimethylcyclopent-3-enyl)ethanal bearing two cyclopropane moieties were synthesized, and the structure of the stereoisomer responsible for its exceptionally strong, diffusive, and natural sandalwood-oil scent, ((1S,2S)-1-methyl-2-{[1S,3R,5R)-1,2,2-trimethylbicyclo[3.1.0]hex-3-yl]methyl}cyclopropyl)methanol (13a), was elucidated.