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 46 毫秒
1.
By a combination of cyclodehydration of N-acyl amino acids with N,N′-diisopropylcarbodiimide (DIC) and non-classical Wittig olefination of the resultant 5(4H)-oxazolones with Ph3PCHCN and Ph3PCHCOOEt, 5-oxazoleacetonitriles and 5-oxazoleacetates were synthesized in one-pot in 41–85% and 57–70% yields, respectively.  相似文献   

2.
The synthesis of two diastereomeric cyclo[Asp-N-Bn-Ser] diketopiperazines (2a and 2b) was investigated. Initial formation of the Boc-aspartyl-N-benzyl serine isopeptide methyl esters (4a and 4b) was observed, which derive from the selective O-acylation of unprotected (S)- or (R)-N-benzylserine. This unexpected O-acylation is preferred over the formation of the tertiary amide and the resulting ester bond is stable in solution to O,N-acyl transfer. The O,N-acyl migration is then triggered by cleavage of the Boc protecting group and treatment with base, which also promotes immediate cyclization to the diketopiperazines.  相似文献   

3.
To better understand reactivity in such systems, fifteen amidoesters derived from β-aminoalcohols were solvolyzed at the ester group in mildly basic methanol-d4. All trials showed pseudo-first-order kinetics by 1H NMR. The rate constants are about 2 to 140-fold larger than those found with simple alkyl esters. The least bulky N-acyl groups generally sponsor the largest rate constants, and strongly so in two cases, but apparently not as a result of lesser steric crowding between the amide and ester groups. Rate constants are also greater for those amidoesters favoring an anti conformation at the amide linkage.  相似文献   

4.
5.
The products of photolysis of N-substituted salicylic acid amides, viz., 2-hydroxy-3-tert-butyl-5-ethylbenzoic acid N-(4-hydroxy-3,5-di-tert-butylphenyl)amide (1) and 2-hydroxybenzoic acid N-[3-(4-hydroxy-3,5-di-tert-butylphenyl)prop-1-yl]amide (2), in heptane were studied by optical spectroscopy and stationary and nanosecond laser photolysis (Nd: YAG laser, 355 nm). It was shown by the method of partial deuteration of amides 1 and 2 that they exist in both the unbound state and as complexes with intraand intermolecular hydrogen bond. Amides 1 and 2 are subjected to photolysis, which results in the formation of a triplet state and phenoxyl radicals RO? presumably due to the absorption of the second photon by the excited singlet state. The formation of radical products due to N–H bond ionization was not observed. The main channel of decay of the triplet state and radicals RO? is triplet–triplet annihilation and recombination (k r ≈ 2.3?108 L mol–1 s–1), respectively. The UV irradiation of compounds 1 and 2 leads to the excitation of the amide groups, and no formation of radical products due to N–H bond ionization was observed.  相似文献   

6.
Anthranilic acid amide reacts with cyclic anhydrides to give the corresponding N-acyl derivatives at the amino group, while analogous reactions of o-aminobenzohydroxamic acid lead lead to formation of 3-hydroxy-quinazolin-4-ones under mild conditions. N-Acyl derivatives of anthranilic acid amide undergo intramolecular cyclization to imides on microwave irradiation or on melting, and their treatment with acetic anhydride in the presence of sodium acetate on heating yields quinazolin-4-ones.  相似文献   

7.
Elena Yu. Schmidt 《Tetrahedron》2009,65(25):4855-4858
2-Arylazo-1-vinylpyrroles 1-6 react with CF3CO2H (benzene, reflux, 5 h) in a peculiar way: instead of expected electrophilic addition of CF3CO2H to the N-vinyl group, the latter is transferred to the azo group followed by NN bond cleavage to afford substituted 2-methylquinolines 10-12 in up to 56% yields. The reaction was shown (1H, 13C, and 15N NMR) to start with the protonation of the azo group with further inter- and intramolecular involving of two protonated N-vinyl groups to finally build up the quinoline cycle over the aryl moiety.  相似文献   

8.
We report the synthesis of N-acetoxy-N-(1-methyl-5H-pyrido[4,5-b]indol-3-yl)acetamide, 7, its N-pivaloyloxy analogue, 9, and improved synthesis of indole-2-acetonitrile, 3 (70% in five steps from indole-2-carboxylic acid), the carcinogenic amine Trp-P-2, 4 (40% from 3), and the nitro compound, 5 (40% from 4 by oxidation with H2O2 using Mo(CO)6 catalyst). In aqueous solution at neutral pH, 7 primarily undergoes C-O bond cleavage to yield the hydroxamic acid, 8, but under the same conditions the sterically hindered 9 decomposes predominately by N-O bond cleavage with a pH independent rate constant that is 7.5-fold smaller than that for 7. In the pH range 0.5-7.0 three different processes for the decomposition of 9 were detected by kinetics. Only the process that dominates at neutral pH generates a nitrenium species that can be trapped by N3.  相似文献   

9.
Michael D. Markey 《Tetrahedron》2008,64(36):8381-8388
With the objective of establishing why reaction of the proposed molecular motors 7 and 22 with carbonyldiimidazole and phosgene does not result in unidirectional rotation, N-ethyl-2-[4-(N,N-dimethylamino)-2-pyridinyl]benzenamine [28, 2-(2-(ethylamino)phenyl)DMAP] was examined as a model substrate. The synthesis of 28 is described. Compound 28 was found to react with phosgene to give the unexpectedly stable N-acyl pyridinium salt 30. The latter (30) is so stable that it is effectively inert to reaction with methanol. At room temperature the two methyls in the Me2N-group of 30 are nonequivalent (NMR) and the barrier to rotation around the Me2N-pyridinium bond is 18.5 kcal/mol. To the authors' knowledge, that is the first quantitative determination of the barrier to rotation around the bond between a 4-(N,N-dimethylamino) group and an N-acyl pyridinium unit. The implications of the findings regarding 30 as to troubleshooting the proposed molecular motor 7, and possible strategies to follow, are discussed.  相似文献   

10.
The N-ferrocenoyl amino acid ester derivatives FcCOR {Fc=(η5-C5H5)Fe(η5-C5H4)} where R=Gly(OMe) 1, Gly(OEt) 2, Gly(OBn) 3, l-Ala(OMe) 4, l-Ala(OEt) 5, l-Leu(OMe) 6, l-Leu(OEt) 7, l-Leu(OBn) 8, l-Phe(OMe) 9 and l-Phe(OEt) 10, were prepared by coupling ferrocene carboxylic acid with the appropriate amino acid ester starting materials using the 1,3-dicyclohexylcarbodiimide (DCC), 1-hydroxybenzotriazole (HOBt) protocol and these have been characterised by spectroscopic techniques. The electrochemical anion sensing behaviour of compounds 1-10 with several anions using a platinum microdisk working electrode is described, together with 1H NMR anion complexation studies. The X-ray single crystal structure of N-ferrocenoyl-l-alanine methyl ester 4 has been determined and contains two molecules which differ slightly in conformation in the asymmetric unit of space group P21 (No. 4); principal dimensions are amide N(H)CO 1.224(6) and 1.231(6) Å, ester CO 1.220(10) and 1.190(7) Å, with N-H?OC(amide) as the primary intermolecular hydrogen bond, N?O 2.992(6) and 2.971(6) Å and with graph set C(4).  相似文献   

11.
Experimentally determined barriers to O-acyl group topomerization in mixed anhydrides composed of β-disubstituted carboxylic acids and cyclic thiohydroxamic acid N-hydroxy-4-methylthiazole-2(3H)-thione were located in the range of ΔG320=68±8 kJ mol−1 (DNMR). According to modeling studies, the underlying exchange process is proposed to occur via rotation about the N,O bond for torsional movement of the O-acyl group past the heterocyclic 4-methyl substituent. The energetically lowest pathway for passing the O-acyl entity by the thione sulfur, is predicted to occur via sequential rocking about the Csp2,O single bond in combination with an interlaced twist about the N,O axis.  相似文献   

12.
Unusual Fragmentation Pathways in Collagen Glycopeptides   总被引:1,自引:0,他引:1  
Collagens are the most abundant glycoproteins in the body. One characteristic of this protein family is that the amino acid sequence consists of repeats of three amino acids –(X—Y—Gly)n. Within this motif, the Y residue is often 4-hydroxyproline (HyP) or 5-hydroxylysine (HyK). Glycosylation in collagen occurs at the 5-OH group in HyK in the form of two glycosides, galactosylhydroxylysine (Gal-HyK) and glucosyl galactosylhydroxylysine (GlcGal-HyK). In collision induced dissociation (CID), collagen tryptic glycopeptides exhibit unexpected gas-phase dissociation behavior compared to typical N- and O-linked glycopeptides (i.e., in addition to glycosidic bond cleavages, extensive cleavages of the amide bonds are observed). The Gal- or GlcGal- glycan modifications are largely retained on the fragment ions. These features enable unambiguous determination of the amino acid sequence of collagen glycopeptides and the location of the glycosylation site. This dissociation pattern was consistent for all analyzed collagen glycopeptides, regardless of their length or amino acid composition, collagen type or tissue. The two fragmentation pathways—amide bond and glycosidic bond cleavage—are highly competitive in collagen tryptic glycopeptides. The number of ionizing protons relative to the number of basic sites (i.e., Arg, Lys, HyK, and N-terminus) is a major driving force of the fragmentation. We present here our experimental results and employ quantum mechanics calculations to understand the factors enhancing the labile character of the amide bonds and the stability of hydroxylysine glycosides in gas phase dissociation of collagen glycopeptides.
Figure  相似文献   

13.
The push-pull characters of a large series of donor-acceptor substituted azo dyes—71 structures in all—have been quantified by the NN double bond lengths, dNN, the 15N NMR chemical shift differences, Δδ15N, of the two nitrogen atoms and the quotient, π/π, of the occupations of the antibonding π, and bonding π orbitals of this partial NN double bond. The excellent correlation of the occupation quotients with the bond lengths strongly infers that both π/π and dNN are excellent parameters for quantifying charge alternation in the push-pull chromophore and the molecular hyperpolarizability, β0, of these compounds. By this approach, selected compounds can be appropriately considered as viable candidates for nonlinear optical (NLO) applications.  相似文献   

14.
Homochiral (E)- and (Z)-enamides derived from SuperQuat (S)-4-phenyl-5,5-dimethyl-oxazolidin-2-one undergo highly diastereoselective epoxidation upon treatment with dimethyldioxirane. Subsequent epoxide opening with meta-chlorobenzoic acid proceeds via a stereoselective SN1-type process, with retention of configuration, to give the corresponding 1′-m-chlorobenzoyl-2′-hydroxy derivatives. Treatment of the SuperQuat enamides with mCPBA effects this two-step transformation in one pot. Reductive cleavage of the isolated 1′-m-chlorobenzoyl-2′-hydroxy derivatives (≥96% de) generates homochiral 1,2-diols in ≥96% ee. Alternatively, regioselective lithiation of the enamide at C(1′) with tBuLi followed by reaction with an aromatic aldehyde and in situ O-benzylation generates a 1′-(benzyloxy-aryl-methyl) substituted enamide with high diastereoselectivity. Subsequent oxidative cleavage of the enamide CC bond with NaIO4/RuCl3 followed by methanolysis of the resultant N-acyl fragment furnishes an O-benzyl protected α-hydroxy methyl ester in high ee.  相似文献   

15.
Complex [RuCl{κ3(N,N,N)-Tp}(PPh3)(PTA)] (κ3(N,N,N)-Tp = hydridotris(pyrazolyl)borate) containing the water-soluble phosphane 1,3,5-triaza-7-phosphatricyclo[3.3.1.13,7]decane (PTA) reacts with terminal alkynes producing to the corresponding neutral alkynyl complexes [Ru(CCR){κ3(N,N,N)-Tp}(PPh3)(PTA)] (R = Ph (1a), nBu (1b), 1-cyclopentenyl (1c), p-methoxyphenyl (1d), 6-methoxynaft-2-yl (1e)). When halide is extracted from complex [RuCl{κ3(N,N,N)-Tp}(PPh3)(PTA)] followed by treatment with propargyl alcohols, the corresponding allenylidene complexes [Ru{κ3(N,N,N)-Tp}(PPh3)(PTA)(CCCPh2)][X] (X = PF6 (2a), CF3SO3 (2b)) and [Ru{κ3(N,N,N)-Tp}(PPh3)(PTA)(CCCC12H8)][PF6] (3) result. Electrophilic attack on the complexes thus obtained leads chemoselectively to the alkynyl complexes [Ru(CCR){κ3(N,N,N)-Tp}(PPh3)(1-CH3-PTA)][CF3SO3] (R = Ph (4a), nBu (4b), and 1-cyclopentenyl (4c)) and to the dicationic allenylidene complexes [Ru{κ3(N,N,N)-Tp}(PPh3)(1-H-PTA)(CCCC12H8)][PF6]2 (5) and [Ru{κ3(N,N,N)-Tp}(PPh3)(1-CH3-PTA)(CCCPh2)][CF3SO3]2 (6).  相似文献   

16.
The inhibition of hydrogen bond formation in the recognition of adipic acid by a new diamide receptor 1 having a pyridine-N-oxide and a simple pyridine ring adjacent to the amide moieties is observed. NMR studies show binding by the pyridine amide group in 1, which demonstrates the discrimination in hydrogen bonding between the carboxyls and an amide adjacent to pyridine versus another adjacent to the pyridine N-oxide. This specific inhibition of hydrogen bonding to a carboxyl group by the two different amides in 1 is corroborated by the NMR binding studies of 1 with propionic acid.  相似文献   

17.
18.
The condensation of (butyl)thiocarbene tungsten complex [(OC)5WC(SEt)Bu] (1a) with an α,β-unsaturated secondary acid amide R2CHCHC(O)NHR14 in the presence of POCl3/Et3N gives cyclopentadienimines 12, whereas the isostructural alkoxycarbene complex [(OC)5WC(OEt)Bu] (1c) under similar conditions affords a (N-enamino)ethoxycarbene compound 9. Furthermore, condensation of the (methyl)thiocarbene tungsten complex [(OC)5WC(SEt)Me] (1b) with an amide 4 yields cyclopentenimines 19 and allenylidene complexes 20, whereas the corresponding ethoxycarbene complex [(OC)5WC(OEt)CH3] (1d) forms 4-NH-amino-1-tungsta-1,3,5-hexatrienes 16 under similar conditions.  相似文献   

19.
Conjugate addition of achiral lithium dimethylamide to the chiral iron cinnamoyl complexes (S,E)- and (S,Z)-[(η5-C5H5)Fe(CO)(PPh3)(COCHCHPh)] proceeds with high diastereoselectivity, with this protocol being used to establish unambiguously the absolute configuration of Winterstein’s acid (3-N,N-dimethylamino-3-phenylpropanoic acid) as (R). The highly diastereoselective conjugate addition of lithium N-benzyl-N-trimethylsilylamide to a range of α,β-unsaturated iron acyl complexes, followed by in-situ elaboration of the derived enolate by either alkylation or aldol reactions is also demonstrated, facilitating the stereoselective synthesis of both cis- and trans-β-lactams. This methodology has been used to effect the formal asymmetric syntheses of (±)-olivanic acid and (±)-thienamycin. Addition of chiral lithium amides derived from primary and secondary amines to the iron crotonyl complex [(η5-C5H5)Fe(CO)(PPh3)(COCHCHMe)] indicates that lithium N-α-methylbenzylamide shows low levels of enantiorecognition, while lithium N-3,4-dimethoxybenzyl-N-α-methylbenzylamide and lithium N-benzyl-N-α-methylbenzylamide show high levels of enantiodiscrimination. The high level of observed enantiorecognition was used to facilitate a kinetic resolution of (RS)-[(η5-C5H5)Fe(CO)(PPh3)(COCHCHMe)] with homochiral lithium (R)-N-3,4-dimethoxybenzyl-N-α-methylbenzylamide. Further mechanistic studies show that conjugate additions of (RS)-lithium N-benzyl-N-α-methylbenzylamide to either the (RS)- or homochiral iron crotonyl complex show 2:1 stoicheiometry, while homochiral lithium N-benzyl-N-α-methylbenzylamide shows 1:1 stoicheiometry.  相似文献   

20.
In the thermolysis of the silaterazolines silatetrazoline tBu2SiNSiCltBu2 · tBu3SiN3 the silanimine tBu2SiNSiCltBu2 and the silyl azide tBu3SiN3 are formed quantitatively. The silanimine tBu2SiNSiCltBu2 has been trapped with Et3NHF, Me3NHCl, water, 1-butene, 2,3-dimethyl-1,3-butadiene, isobutene, methylvinyl ether, and tBu2SiClN3. The structure of the disiloxane (tBu2SiCl-NH-SitBu2)2O and of the bis(di-tert-butylchlorsilyl)-substituted silatetrazoline tBu2SiNSiCltBu2 · tBu2SiClN3 has been determined by X-ray structure analysis.  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