首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
In the course of the first of several attempts to elaborate methods for the synthesis of 1‐nitropiperidinoses, lincosamine was transformed into lactam 6 via hemiacetal 1 , lactone 2 , amide 3 , oxo amide 4 , and its cyclic tautomer 5 . Treatment of the N‐Boc‐protected lactam oxime 9 , obtained from lactam 6 , with brominating agents failed to provide the bromonitroso carbamate 10 . The N‐Boc‐protected lactam 13 derived from 6 was reduced to hemiacetal 14 , but the corresponding N‐Boc‐aminooxime did not tautomerise to the C(1)‐hydroxylamine, and nitrone 17 , a potential precursor of the nitropiperidine 12 , was not formed. Oxidation of the anomeric azide 20 with HOF?MeCN failed to provide the expected nitropiperidine 21 . The phosphinimines 22 derived from 20 did not react with O3. In the next approach to 1‐nitropiperidinoses, we treated the N‐Boc‐protected hemiacetal 25 , obtained from the known gluconolactam 23 with N‐benzylhydroxylamine. The resulting nitrone 26 exits in equilibrium with the anomeric N‐benzyl‐glycosylhydroxylamine that was oxidized to the anomeric nitrone 28 . Ozonolysis of 28 led to the hemiacetal 25 , resulting from the desired, highly reactive protected nitropiperidinose 29 , that was evidenced by an IR band at 1561 cm?1. Similarly to the synthesis of nitrone 26 , reaction of the N‐tosyl‐protected hemiacetal 31 with N‐benzylhydroxylamine and oxidation provided the anomeric N‐benzylhydroxylamines 33 via the p‐toluenesulfonamido nitrone 32 . Their oxidation with MnO2 led to the anomeric nitrone 34 . Ozonolysis of 34 as evidenced by 1H‐NMR and ReactIR spectroscopy led to the highly reactive nitropiperidinose 35 . Like 29, 35 was transformed during workup, and only the hemiacetal 31 was isolated. The similarly prepared lincosamine‐derived nitrone 17 was subjected to ReactIR‐monitored ozonolysis that evidenced the formation of the protected nitropiperidinose 12 , but only led to the isolation of 14 . The facile transformation of the nitropiperidinoses to hemiacetals is rationalised by heterolysis of the anomeric C,N bond, recombination of the ion pair, and denitrosation of the resulting anomeric nitrite by a nucleophile. Attempts to convert the 1‐deoxy‐1‐nitropiperidinose 35 to uloses 43 by base‐catalysed Michael additions or Henry reactions were unsuccessful.  相似文献   

2.
α‐Fluorinated β‐amino thioesters were obtained in high yields and stereoselectivities by organocatalyzed addition reactions of α‐fluorinated monothiomalonates (F‐MTMs) to N‐Cbz‐ and N‐Boc‐protected imines. The transformation requires catalyst loadings of only 1 mol % and proceeds under mild reaction conditions. The obtained addition products were readily used for coupling‐reagent‐free peptide synthesis in solution and on solid phase. The α‐fluoro‐β‐(carb)amido moiety showed distinct conformational preferences, as determined by crystal structure and NMR spectroscopic analysis.  相似文献   

3.
A one‐pot liquid‐phase combinatorial synthesis of 2‐(4‐oxo‐4H‐1‐benzopyran‐3‐yl)‐4‐thiazolidinones bearing diverse substituents at the 3‐position under microwave irradiation was successfully performed using 3‐formyl chromone, primary amine, and mercaptoacetic acid as reactants. Compared to an identical library generated by conventional parallel synthesis, the microwave‐assisted parallel synthesis approach dramatically decreased the reaction time from an average of 9 h to 5 min, and substantially increased the product yields. The coupling of microwave technology with liquid‐phase combinatorial synthesis constitutes a novel and particularly attractive avenue for the rapid generation of structurally diverse libraries. © 2007 Wiley Periodicals, Inc. Heteroatom Chem 18:381–389, 2007; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20309  相似文献   

4.
Combinatorial diversity in hypervariable β‐hairpin loops is exploited by the immune system to select binding sites on antibodies for a wide variety of different protein antigens. In a first step towards mimicking this strategy in vitro, for the selection of novel protein ligands, an approach is described here for the parallel synthesis of small libraries of conformationally defined β‐hairpin protein epitope mimetics. Starting from a protruding hairpin loop in platelet‐derived growth factor B (PDGF‐B), 8 and 12 residues were first transplanted from the protein to a D ‐Pro‐L ‐Pro template, to afford the cyclic peptide‐loop mimetics 1 and 2 , respectively. NMR and MD studies in aqueous solution show that both mimetics populate conformations which closely mimic the β‐hairpin in the crystal structure of the native protein (Fig. 5). Based on 1 as a scaffold, a library of 24 mimetics was synthesized in which the four residues at the tip of the loop (VRKK) were held constant, and flanking residues at positions 1, 2, 7, and 8 in the hairpin were varied (Fig. 7). The library was prepared by parallel synthesis in a two‐stage solid‐phase assembly/solution‐phase cyclization process. The products were analyzed by MS, NMR, and CD. 2D‐NOESY revealed for most library members characteristic long‐range NOEs that show that the hairpin conformation is stably maintained. The results suggest that this approach may be useful for the synthesis of much larger libraries of peptide and protein mimetics based on a β‐hairpin scaffold.  相似文献   

5.
The synthesis of several Ntert‐butoxycarbonyl(Boc)‐protected‐N‐substituted hydrazines has been accomplished. The use of these protected hydrazines in SNAr substitutions leads to products in which the most nucleophilic nitrogen displaces the leaving group. Treatment of these compounds with trifluoroacetic acid readily removes the Boc‐protecting group and the intermediates readily undergo cyclizations to yield N‐1‐substituted aza‐benzothiopyranoindazoles, anthrapyrazoles and aza‐anthrapyrazoles. Side chain buildup was employed in the synthesis of several aza‐anthrapyrazoles.  相似文献   

6.
A highly stereoselective and efficient total synthesis of trans‐dihydronarciclasine from a readily available chiral starting material was developed. The synthesis defines two of the five stereogenic centers of the natural product by an amino acid ester–enolate Claisen rearrangement. The other three stereogenic centers are created in a highly stereocontrolled fashion via a six‐ring vinylogous ester intermediate, which is generated from the γ,δ‐unsaturated ester functional group of the Claisen rearrangement product in an efficient three‐step sequence. This concise total synthesis exemplifies the use of a highly regioselective Friedel–Crafts‐type cyclization to form the B ring via an isocyanate intermediate derived from an N‐Boc group, which is superior to the conventional method using an imino triflate intermediate. This same N‐Boc group is employed to give high selectivity in the Claisen rearrangement earlier in the sequence.  相似文献   

7.
4‐(4,6‐Dimethoxy‐1,3,5‐triazin‐2‐yl)‐4‐methylmorpholinium toluene‐4‐sulfonate (DMT/NMM/TsO), a representative member of the inexpensive and environmentally‐friendly N‐triazinylammonium family of sulfonates, has been found to be a very effective coupling reagent for the synthesis of amides, esters and peptides in solution. This study confirms the usefulness of DMT/NMM/TsO for peptide synthesis in solution, starting from Z‐, Fmoc‐, and Boc‐protected substrates as well as unnatural building blocks. Peptide synthesis with DMT/NMM/TsO produced high yields, with high crude product purity and low risk of racemization. In all cases, stoichiometric amounts of reagents were used and the standard synthetic procedure, without the need for time‐consuming optimization stages or expensive chromatographic purification. DMT/NMM/TsO was also found to be very useful for the synthesis of oligopeptides using a fragment coupling strategy.  相似文献   

8.
A flexible approach to ethyl (3R,4S)-N-Boc-4-amino-3-hydroxy-5-phenylpentanoate (N-Boc-AHPPA-OEt), the γ-amino-β-hydroxy acid moiety of hapalosin is described. The synthetic method features a ring-opening ethanolysis of an activated N-Boc-lactam, which is obtained via a diastereoselective reductive-alkylation of (R)-malimide derivative. The flexibility of the method resides in the introduction of the alkyl side chain by Grignard reagent addition.  相似文献   

9.
Diastereoselective C3‐elongation processes of N‐Boc‐, N‐Z‐, N‐Bn‐N‐Boc‐, and N‐Bn‐N‐Z‐L ‐alaninals (Boc=tBuOCO, Z=PhCH2OCO, Bn=PhCH2) using various allyl reagents, such as allyl bromide in the presence of Zn/aqueous NH4Cl solution, of SnCl2⋅2 H2O/NaI or of Mg/CuCl2⋅2 H2O, as well as allyltrichlorosilane, are described. A substantially different influence of the N‐protecting groups replacing either one or two amino protons was observed, allowing the selective synthesis of either the syn‐ or anti‐diastereoisomer as a major product.  相似文献   

10.
Five sets of 27‐membered combinatorial libraries of alicyclic β‐lactams were prepared via liquid‐phase Ugi 4‐center 3‐component reactions (U‐4C‐3CR) utilizing 3 different cis β‐amino acids, 3 different isonitriles and 5×3 sets of aldehydes. Through combinations of the building blocks of one of these libraries, all of the possible sublibraries were also generated. A few azetidinone derivatives were synthesized individually by parallel synthesis.  相似文献   

11.
β‐Peptides offer the unique possibility to incorporate additional heteroatoms into the peptidic backbone (Figs. 1 and 2). We report here the synthesis and spectroscopic investigations of β2‐peptide analogs consisting of (S)‐3‐aza‐β‐amino acids carrying the side chains of Val, Ala, and Leu. The hydrazino carboxylic acids were prepared by a known method: Boc amidation of the corresponding N‐benzyl‐L ‐α‐amino acids with an oxaziridine (Scheme 1). Couplings and fragment coupling of the 3‐benzylaza‐β2‐amino acids and a corresponding tripeptide (N‐Boc/C‐OMe strategy) with common peptide‐coupling reagents in solution led to β2‐di, β2‐tri‐, and β2‐hexaazapeptide derivatives, which could be N‐debenzylated ( 4 – 9 ; Schemes 2–4). The new compounds were identified by optical rotation, and IR, 1H‐ and 13C‐NMR, and CD spectroscopy (Figs. 4 and 5) and high‐resolution mass spectrometry, and, in one case, by X‐ray crystallography (Fig. 3). In spite of extensive measurements under various conditions (temperatures, solvents), it was not possible to determine the secondary structure of the β2‐azapeptides by NMR spectroscopy (overlapping and broad signals, fast exchange between the two types of NH protons!). The CD spectra of the N‐Boc and C‐OMe terminally protected hexapeptide analog 9 in MeOH and in H2O (at different pH) might arise from a (P)‐314‐helical structure. The N‐Boc‐β2‐tri and N‐Boc‐β2‐hexaazapeptide esters, 7 and 9 , were shown to be stable for 48 h against the following peptidases: pronase, proteinase K, chymotrypsin, trypsin, carboxypeptidase A, and 20S proteasome.  相似文献   

12.
A highly stereoselective aza‐Henry reaction of α‐aryl nitromethanes with aromatic N‐Boc imines was established by using C1‐symmetric chiral ammonium betaine as a bifunctional organic base catalyst. Various substituted aryl groups for both imines and nitromethanes were tolerated in the reaction, and a series of precursors for the synthesis of unsymmetrical anti‐1,2‐diaryl ethylenediamines was provided.  相似文献   

13.
N‐Boc/Fmoc/Z‐N′‐formyl‐gem‐diaminoalkyl derivatives, intermediates particularly useful in the synthesis of partially modified retro‐inverso peptides, have been characterized by both positive and negative ion electrospray ionization (ESI) ion‐trap multi‐stage mass spectrometry (MSn). The MS2 collision induced dissociation (CID) spectra of the sodium adduct of the formamides derived from the corresponding N‐Fmoc/Z‐amino acids, dipeptide and tripeptide acids show the [M + Na‐NH2CHO]+ ion, arising from the loss of formamide, as the base peak. Differently, the MS2 CID spectra of [M + Na]+ ion of all the N‐Boc derivatives yield the abundant [M + Na‐C4H8]+ and [M + Na‐Boc + H]+ ions because of the loss of isobutylene and CO2 from the Boc protecting function. Useful information on the type of amino acids and their sequence in the N‐protected dipeptidyl and tripeptidyl‐N′‐formamides is provided by MS2 and subsequent MSn experiments on the respective precursor ions. The negative ion ESI mass spectra of these oligomers show, in addition to [M‐H]?, [M + HCOO]? and [M + Cl]? ions, the presence of in‐source CID fragment ions deriving from the involvement of the N‐protecting group. Furthermore, MSn spectra of [M + Cl]? ion of N‐protected dipeptide and tripeptide derivatives show characteristic fragmentations that are useful for determining the nature of the C‐terminal gem‐diamino residue. The present paper represents an initial attempt to study the ESI‐MS behavior of these important intermediates and lays the groundwork for structural‐based studies on more complex partially modified retro‐inverso peptides. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

14.
The first total synthesis of sphingolipid (2S,3R,4E)‐N2‐octadecanoyl‐4‐tetradecasphingenine ( 1a ), a natural sphingolipid isolated from Bombycis Corpus 101A, and of its styryl analogue 1b was achieved in good overall yield (Schemes 1 and 2). The key step involved the installation with (E) stereoselectivity of a long lipophilic chain or phenyl group on allyl alcohol derivative 3 via a cross‐metathesis reaction (→ 5a or 5b ). The N‐Boc protected 3 was easily accessible from (S)‐Garner aldehyde.  相似文献   

15.
The known solid‐state structure (Fig. 1, top) of cyclo(β‐HAla)4 was used to model the structure of the title compound 1 as a prospective somatostatin mimic (Fig. 1, bottom). The synthesis started with the N‐protected natural amino acids Boc‐Phe‐OH, Boc‐Trp‐OH, Boc‐Lys(2‐Cl‐Z)‐OH, and Boc‐Thr(OBn)‐OH, which were homologated to the corresponding β‐amino‐acid derivatives (Scheme 1) and coupled to the β‐tetrapeptide Boc‐β‐HTrp‐β‐HPhe‐β‐HThr(OBn)‐β‐HLys(2‐Cl‐Z)‐OMe ( 16 ); the (N‐Me)‐β‐HThr‐(N‐Me)‐β‐HPhe analog 17 was also prepared. C‐ and N‐terminal deprotection and cyclization through the pentafluorophenyl ester gave the insoluble β‐tetrapeptide with protected Thr and Lys side chains ( 18 ). Solubilization and debenzylation could only be effected in LiCl‐containing THF (ca. 10% yield; with ca. 55% recovery). HPLC Purification provided a sample of the title compound 1 , the structure of which, as determined by NMR‐spectroscopy (Fig. 2, left) was drastically different from the `theoretical' model (Fig. 1). There is a transannular H‐bond dividing the macrocyclic 16‐membered ring, thus forming a ten‐ and a twelve‐membered H‐bonded ring, the former mimicking, or actually being superimposable on, an α‐peptidic so‐called β‐turn. Still, the four side chains occupy equatorial positions on the ring, as planned, albeit with somewhat different geometry as compared to the `original'. The cycloβ‐tetrapeptide has micromolar affinities to the human somatostatin receptors (hsst 1 – 5). Thus, we have demonstrated for the first time that it is possible to mimic a natural peptide hormone with a small β‐peptide. Furthermore, we have discovered a simple way to construct the ubiquitous β‐turn motif with β‐peptides (which are known to be stable to mammalian peptidases).  相似文献   

16.
The title compound (systematic name: methyl 2‐{2‐[(tert‐butoxycarbonyl)amino]‐2‐methylpropanamido}‐2‐methylpropanoate), C14H26N2O5, (I), crystallizes in the monoclinic space group P21/n in two polymorphic forms, each with one molecule in the asymmetric unit. The molecular conformation is essentially the same in both polymorphs, with the α‐aminoisobutyric acid (Aib) residues adopting ϕ and ψ values characteristic of α‐helical and mixed 310‐ and α‐helical conformations. The helical handedness of the C‐terminal residue (Aib2) is opposite to that of the N‐terminal residue (Aib1). In contrast to (I), the closely related peptide Boc‐Aib‐Aib‐OBn (Boc is tert‐butoxycarbonyl and Bn is benzyl) adopts an αL‐PII backbone conformation (or the mirror image conformation). Compound (I) forms hydrogen‐bonded parallel β‐sheet‐like tapes, with the carbonyl groups of Aib1 and Aib2 acting as hydrogen‐bond acceptors. This seems to represent an unusual packing for a protected dipeptide containing at least one α,α‐disubstituted residue.  相似文献   

17.
2‐Thioxo‐5,7‐dimethylpyrido[2,3‐d]pyrimidin‐4(3H)‐ones 3 were synthesized by the cyclocondensation of 2‐amino‐3‐carbethoxy‐4,6‐dimethylpyridine 1 with methyl‐N‐aryldithiocarbamates 2 and compared with the condensation between 1 and aryl isothiocyanates 4. When a comparative study of N vs S alkylation of ambident 2‐thioxo‐5,7‐dimethylpyrido[2,3‐d]pyrimidin‐4(3H)‐ones 3 was carried out under liquid‐liquid and solid‐liquid phase transfer conditions using various alkylating agents 5 , the S‐alkylated products 6 were obtained exclusively and selectively.  相似文献   

18.
A chiral bis(guanidino)iminophosphorane catalyzes enantioselective addition reactions of a 1,3‐dithiane derivative as a pronucleophile. The chiral uncharged organosuperbase facilitates the addition of benzyloxycarbonyl‐1,3‐dithiane to aromatic N‐Boc‐protected imines to provide optically active α‐amino‐1,3‐dithiane derivatives, which are valuable versatile building blocks in organic synthesis.  相似文献   

19.
A new diastereoselective route to 2‐aminotetrahydrofurans has been developed from N,O‐dialkenylhydroxylamines. These intermediates undergo a spontaneous C?C bond‐forming [3,3]‐sigmatropic rearrangement followed by a C?O bond‐forming cyclization. A copper‐catalyzed N‐alkenylation of an N‐Boc‐hydroxylamine with alkenyl iodides, and a base‐promoted addition of the resulting N‐hydroxyenamines to an electron‐deficient allene, provide modular access to these novel rearrangement precursors. The scope of this de novo synthesis of simple nucleoside analogues has been explored to reveal trends in diastereoselectivity and reactivity. In addition, a base‐promoted ring‐opening and Mannich reaction has been discovered to covert 2‐aminotetrahydrofurans to cyclopentyl β‐aminoacid derivatives or cyclopentenones.  相似文献   

20.
Proton abstraction of Ntert‐butoxycarbonyl‐piperidine (N‐Boc‐piperidine) with sBuLi and TMEDA provides a racemic organolithium that can be resolved using a chiral ligand. The enantiomeric organolithiums can interconvert so that a dynamic resolution occurs. Two mechanisms for promoting enantioselectivity in the products are possible. Slow addition of an electrophile such as trimethylsilyl chloride allows dynamic resolution under kinetic control (DKR). This process occurs with high enantioselectivity and is successful by catalysis with substoichiometric chiral ligand (catalytic dynamic kinetic resolution). Alternatively, the two enantiomers of this organolithium can be resolved under thermodynamic control with good enantioselectivity (dynamic thermodynamic resolution, DTR). The best ligands found are based on chiral diamino‐alkoxides. Using DTR, a variety of electrophiles can be used to provide an asymmetric synthesis of enantiomerically enriched 2‐substituted piperidines, including (after Boc deprotection) the alkaloid (+)‐β‐conhydrine. The chemistry was extended, albeit with lower yields, to the corresponding 2‐substituted seven‐membered azepine ring derivatives.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号