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.
The stabilization of peptide secondary structure via stapling is a ubiquitous goal for creating new probes, imaging agents, and drugs. Inspired by indole‐derived crosslinks found in natural peptide toxins, we employed ortho‐phthalaldehydes to create isoindole staples, thus transforming inactive linear and monocyclic precursors into bioactive monocyclic and bicyclic products. Mild, metal‐free conditions give an array of macrocyclic α‐melanocyte‐stimulating hormone (α‐MSH) derivatives, of which several isoindole‐stapled α‐MSH analogues (Ki≈1 nm ) are found to be as potent as α‐MSH. Analogously, late‐stage intra‐annular isoindole stapling furnished a bicyclic peptide mimic of α‐amanitin that is cytotoxic to CHO cells (IC50=70 μm ). Given its user‐friendliness, we have termed this approach FlICk (fluorescent isoindole crosslink) chemistry.  相似文献   

2.
Xiao Hu  Bo Wu  Weiping Zheng 《中国化学》2019,37(3):244-248
Even though the bis‐lactam peptide stapling with the [i, i + 7] and the [i, i + 11] systems has been known to be able to afford % α‐helicity values up to 100% (25°C), the performance of the bis‐lactam peptide stapling with the [i, i + 4] system in current literature has been mediocre (% α‐helicity ≦40%, 25°C). In the current study, we found that high % α‐helicity is also obtainable with the bis‐lactam [i, i + 4]‐stapling by demonstrating with our model peptide sequence that the bis‐lactam [i, i + 4]‐stapling with Nεpara‐phenylenediacetyl‐lysine was able to afford a % α‐helicity value of ~64.1% (25°C). Therefore, the bis‐lactam [i, i + 4]‐stapling could also be an efficacious peptide stapling mode that can be employed for biomedical applications.  相似文献   

3.
The organocatalytic properties of unnatural α‐amino acids are reviewed. Post‐translational derivatives of natural α‐amino acids include 4‐hydroxy‐l ‐proline and 4‐amino‐l ‐proline scaffolds, and also proline homologues. The activity of synthetic unnatural α‐amino acid‐based organocatalysts, such as β‐alkyl alanines, alanine‐based phosphines, and tert‐leucine derivatives, are reviewed herein. The organocatalytic properties of unnatural monocyclic, bicyclic, and tricyclic proline derivatives are also reviewed. Several families of these organocatalysts permit the efficient and stereoselective synthesis of complex natural products. Most of the reviewed organocatalysts accelerate the reported reactions through covalent interactions that raise the HOMO (enamine intermediates) or lower the LUMO (iminium intermediates).  相似文献   

4.
The stereoselective syntheses of 7,8,9‐trideoxypeloruside A ( 4 ) and a monocyclic peloruside A analogue lacking the entire tetrahydropyran moiety ( 3 ) are described. The syntheses proceeded through the PMB‐ether of an ω‐hydroxy β‐keto aldehyde as a common intermediate which was elaborated into a pair of diastereomeric 1,3‐syn and ‐anti diols by stereoselective Duthaler–Hafner allylations and subsequent 1,3‐syn or anti reduction. One of these isomers was further converted into a tetrahydropyran derivative in a high‐yielding Prins reaction, to provide the precursor for bicyclic analogue 4 . Downstream steps for both syntheses included the substrate‐controlled addition of a vinyl lithium intermediate to an aldehyde, thus connecting the peloruside side chain to C15 (C13) of the macrocyclic core structure in a fully stereoselective fashion. In the case of monocyclic 3 macrocyclization was based on ring‐closing olefin metathesis (RCM), while bicyclic 4 was cyclized through Yamaguchi‐type macrolactonization. The macrolactonization step was surprisingly difficult and was accompanied by extensive cyclic dimer formation. Peloruside A analogues 3 and 4 inhibited the proliferation of human cancer cell lines in vitro with micromolar and sub‐micromolar IC50 values, respectively. The higher potency of 4 highlights the importance of the bicyclic core structure of peloruside A for nM biological activity.  相似文献   

5.
The toxic bicyclic octapeptide α‐amanitin is mostly found in different species of the mushroom genus Amanita, with the death cap (Amanita phalloides) as one of the most prominent members. Due to its high selective inhibition of RNA polymerase II, which is directly linked to its high toxicity, particularly to hepatocytes, α‐amanitin received an increased attention as a toxin‐component of antibody‐drug conjugates (ADC) in cancer research. Furthermore, the isolation of α‐amanitin from mushrooms as the sole source severely restricts compound supply as well as further investigations, as structure–activity relationship (SAR) studies. Based on a straightforward access to the non‐proteinogenic amino acid dihydroxyisoleucine, we herein present a robust total synthesis of α‐amanitin providing options for production at larger scale as well as future structural diversifications.  相似文献   

6.
In this report, we have synthesized organic/inorganic hybrid peptide–poly(?‐caprolactone) (PCL) conjugates via ring opening polymerization (ROP) of ?‐caprolactone (CL) in the presence of two sequence defined peptide initiators, namely POSS‐Leu‐Aib‐Leu‐NH2 (POSS: polyhedral oligomeric silsesquioxane; Leu: Leucine; Aib: α‐aminoisobutyric acid) and OMe‐Leu‐Aib‐Leu‐NH2. Covalent attachment of peptide segments with the PCLs were examined by 1H and 29Si NMR spectroscopy, matrix‐assisted laser desorption/ionization time‐of‐flight mass spectrometry (MALDI‐TOF‐MS) and FTIR spectroscopy. Supramolecular inclusion complexations of synthesized peptide‐PCL conjugates with α‐cyclodextrin (α‐CyD) were studied to understand the effect of POSS/OMe‐peptide moieties at the PCL chain ends. Inclusion complexation of peptide‐PCL conjugates with α‐CyD produced linear polypseudorotaxane, confirmed by 1H NMR, FTIR, powder X‐ray diffraction (PXRD), polarized optical microscopy (POM) and differential scanning calorimetry (DSC). Extent of α‐CyD threading onto the hybrid peptide‐PCL conjugated polymers is less than that of α‐CyD threaded onto the linear PCL. Thus, PCL chains were not fully covered by the host α‐CyD molecules due to the bulky POSS/OMe‐peptide moieties connected with the one edge of the PCL chains. PXRD experiment reveals channel like structures by the synthesized inclusion complexes (ICs). Spherulitic morphologies of POSS/OMe‐peptide‐PCL conjugates were fully destroyed after inclusion complexation with α‐CyD and tiny nanoobjects were produced. © 2016 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 3643–3651.  相似文献   

7.
Thiopeptides, arising from complex posttranslational modifications of a genetically encoded precursor peptide, are of great interest due to their structural complexity and important biological activities. All of these antibiotics share a macrocyclic peptidyl core that contains a central, six‐membered nitrogen heterocycle and are classified into five series ae based on the oxidation state of the central nitrogenous ring. Here, we report that the biosynthesis of the central piperidine heterocycle of series a thiopeptides relies on the activity of homologues of an F420H2‐dependent reductase TppX4 by exploiting and characterizing the piperidine‐containing thiopeptin biosynthetic gene (tpp) cluster in Streptomyces tateyamensis. In vitro reconstruction of TppX4 activity demonstrated that the piperidine heterocycle of thiopeptins was transformed from a dehydropiperidine heterocycle, and TppX4 tolerated the changes in the C‐termini and macrocyclic peptidyl core of substrate and also tolerated dehyropiperidine‐containing monocyclic or bicyclic thiopeptides. The identification of TppX4 and its substrate tolerance enriches the biosynthetic toolbox for development of additional thiopeptide analogs for clinical drug screening.  相似文献   

8.
The ring‐collapse mechanism of C24 (D6d) has been analyzed using semiempirical AM1 and B3LYP/cc‐pVDZ methods. Based on the ring‐stacking/circumscribing model, two precursors are selected. Transition states and intermediates are located and energetics are computed. Before the stacking begins, the precursor and belt reach a suitable relative orientation accompanied by the release of a large amount of energy. It is observed that the reactions between the precursors and the belts are essentially endoergic in nature, whereas the reactions between the stable intermediates and the final belts are exoergic. The deformation energies (DE) and the bond lengths R of the precursors have been computed. The DE values suggest that there is a chance of the cleavage of the bicyclic precursors as the growth process proceeds toward the cage formation. In contrast, the monocyclic precursor is found to have lower deformation energies than the bicyclic precursor. Analysis of average bond length at different cages shows that a large window is formed and the system appears to follow a cascade‐type bond formation. Comparisons are made to our previous results on C28 growth. © 2005 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

9.
Debromination of N‐benzyl 4‐ or 5‐substituted α‐bromo‐α‐p‐toluenesulfonylglutarimides is achieved with 1,8‐diazabicyclo[5.4.0]undec‐7‐ene (DBU) to give the N‐benzyl 4‐ or 5‐substituted α‐p‐toluenesulfonylglutarimides. The DBU/THF system is applied to a new methodology for the synthesis of bicyclic glutarimide skeleton in moderate yields.  相似文献   

10.
Maintaining specific conformations of peptide ligands is crucial for improving the efficacy of biological interactions. Here, a one‐pot polymerization strategy for stabilizing the α‐helical conformation of peptides while simultaneously constructing multimeric ligands is presented. The new method, termed stapling polymerization, uses radical polymerization between acryloylated peptide side chains and vinylic monomers. Studies with model peptides indicate that i, i+7 crosslinking is effective for the helix stabilization, whereas i, i+4 crosslinking is not. The stapling polymerization results in the formation of peptide–polyacrylamide conjugates that include ≈3–16 peptides in a single conjugate. This stapling polymerization provides a simple but powerful methodology to fabricate multimeric α‐helices that can further be developed to modulate multivalent biomacromolecular interactions.

  相似文献   


11.
Reactions of bicyclic α‐P4S3I2 with Hpthiq gave solutions containing α‐P4S3(pthiq)I and α‐P4S3(pthiq)2, where Hpthiq is the conformationally constrained chiral secondary amine 1‐phenyl‐1,2,3,4‐tetrahydroisoquinoline. The expected diastereomers have been characterised by complete analysis of their 31P{1H} NMR spectra. Hindered P–N bond rotation in the amide iodide α‐P4S3(pthiq)I caused greater broadening of peaks in the room‐temperature spectrum of one diastereomer than in that of the other. At 183 K, spectra of two P–N bond rotamers for each diastereomer were observed and analysed. The minor rotamers showed strong evidence for steric crowding, having large diastereomeric differences in 1J(P–P) and 2J(P–S–P) couplings (49 Hz, 16 % of value, and 4.4 Hz, 19 % of value, respectively).  相似文献   

12.
The synthesis of α‐sialosides is one of the most difficult reactions in carbohydrate chemistry and is considered to be both a thermodynamically and kinetically disfavored process. The use of acetonitrile as a solvent is an effective solution for the α‐selective glycosidation of N‐acetyl sialic acids. In this report, we report on the α‐glycosidation of partially unprotected N‐acetyl and N‐glycolyl donors in the absence of a nitrile solvent effect. The 9‐O‐benzyl‐N‐acetylthiosialoside underwent glycosidation in CH2Cl2 with a good α‐selectivity. On the other hand, the 4,7,8‐O‐triacetyl‐9‐O‐benzyl‐N‐acetylthiosialoside was converted to β‐sialoside as a major product under the same reaction conditions. The results indicate that the O‐acetyl protection of the sialyl donor was a major factor in reducing the α‐selectivity of sialylation. After tuning of the protecting groups of the hydroxy groups at the 4,7,8 position on the sialyl donor, we found that the 9‐O‐benzyl‐4‐O‐chloroacetyl‐N‐acetylthiosialoside underwent sialylation with excellent α‐selectivity in CH2Cl2. To demonstrate the utility of the method, straightforward synthesis of α(2,9) disialosides containing N‐acetyl and/or N‐glycolyl groups was achieved by using the two N‐acetyl and N‐glycolyl sialyl donors.  相似文献   

13.
In recent years, the strategy of inhibiting the interactions of p53 with murine double minute 2(MDM2)and murine double minute X(MDMX) has been proved to be a promising approach for tumor therapy.However, the poor proteolytical stability and low intracellular delivery efficiency of peptide inhibitors limit their clinical application. Here, we designed and synthesized the bicyclic stapled peptides based on p53 by combining all-hydrocarbon stapling and lactam stapling strategies. We demonstrated th...  相似文献   

14.
The molecular structures of 2‐[4‐chloro‐2‐fluoro‐5‐(prop‐2‐ynyloxy)phenyl]‐1,3,4,5,6,7‐hexahydro­isoindole‐1,3‐dione, C17H13ClFNO3, (I), and the isomeric compound 2‐[4‐chloro‐2‐fluoro‐5‐(prop‐2‐ynyloxy)phenyl]‐cis‐1,3,3a,4,7,7a‐hexahydro­isoindole‐1,3‐dione, (II), are, as anticipated, significantly different in their conformations and in the distances between the farthest two atoms. The six‐membered ring of the 1,3,4,5,6,7‐hexahydro­isoindole‐1,3‐dione moiety in (I) adopts a half‐chair conformation. The dihedral angle between the five‐membered dione ring of (I) and the benzene ring is 50.96 (7)°. The six‐membered ring of the cis‐1,3,3a,4,7,7a‐hexahydro­isoindole‐1,3‐dione moiety in (II) adopts a boat conformation. The dihedral angle in (II) between the five‐membered dione ring and the benzene ring is 61.03 (13)°. In the crystal structures, the molecules are linked by C—H⋯O hydrogen bonds and weak π–π interactions. Compound (I) is a much more potent herbicide than (II). The Cl⋯H distances between the farthest two atoms in (I) and (II) are 11.37 and 9.97 Å, respectively.  相似文献   

15.
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   

16.
The three‐component reaction of the zwitterions generated from dialkyl acetylenedicarboxylates (=dialkyl but‐2‐ynedioates and triphenylphosphine (Ph3P) with isoindoline‐1,3‐diimine (=1H‐isoindole‐1,3(2H)‐diimine) is described (Scheme 1). This reaction affords the corresponding special type of substituted dihydropyrimido[2,1‐a]isoindole derivatives in good yields without using any catalyst and activation (Table).  相似文献   

17.
Reaction of a mixture of bicyclic phosphorus sulfide selenide iodides α‐P4SnSe3−nI2 (n = 0–3) with PriNH2 and Et3N gave corresponding diamides α‐P4SnSe3−n(NHPri)2 (n = 0–3) and imides α‐P4SnSe3−n(μ‐NPri) (n = 2–3), identified in solution by 31P NMR. In one isomer of α‐P4S2Se(μ‐NPri), the C2 symmetry of imides such as α‐P4S3(μ‐NPri) was broken, allowing relative assignment of 2J NMR couplings to the PNP bridge and the PSP bridge opposite to it. The coupling through the sulfur bridge was found to be reduced to ca. zero, in contrast to previous assumptions for this class of compounds. Ab initio models were calculated at the MPW1PW91/svp level for the sulfide selenide imides and for a selection of bond rotamers of the diamides, and at the MPW1PW91/LanL2DZ(d) level for the sulfide selenide diiodides. Different skeletal isomers were prevalent for the mixed chalcogenide diamides than for the diiodides, showing that exchange of chalcogen between skeletal positions took place in the amination reaction even at room temperature. Similar differences to those observed were predicted by the models, suggesting that equilibrium was attained.  相似文献   

18.
Enantioselective α‐arylation of racemic ketones containing existing α′‐stereocenters is reported for the first time via base‐induced dynamic kinetic resolution. Bridged bicyclic rings are formed in good ee values, which are difficult to obtain otherwise. Furthermore, reactions in DMSO‐d6 resulted in extensive deuteration of both α‐ and α′‐positions in the products, thus supporting a pathway involving rapid, reversible deprotonation of ketones under catalytic conditions.  相似文献   

19.
The small synthetic peptide, benzyl 2‐(tert‐but­oxy­carbonyl‐amino)­isobutyrate, C16H23NO4, has the α‐helical conformation [|?| = 55.8 (2)° and |ψ| = 37.9 (2)°] observed in peptide fragments of peptaibols containing the α‐amino­isobutyric acid (Aib) residue. The structure shows no intramolecular hydrogen bonding, which would disrupt the limited conformational freedom associated with this amino acid. Two weak intermolecular hydrogen contacts are observed.  相似文献   

20.
“Stapled” peptides are typically designed to replace two non‐interacting residues with a constraining, olefinic staple. To mimic interacting leucine and isoleucine residues, we have created new amino acids that incorporate a methyl group in the γ‐position of the stapling amino acid S5. We have incorporated them into a sequence derived from steroid receptor coactivator 2, which interacts with estrogen receptor α. The best peptide (IC50=89 nm ) replaces isoleucine 689 with an S‐γ‐methyl stapled amino acid, and has significantly higher affinity than unsubstituted peptides (390 and 760 nm ). Through X‐ray crystallography and molecular dynamics studies, we show that the conformation taken up by the S‐γ‐methyl peptide minimizes the syn‐pentane interactions between the α‐ and γ‐methyl groups.  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