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.
Abstract: The chemical synthesis and X‐ray crystal structure analysis of a model peptide incorporating a conformationally flexible β‐Ala residue: Boc‐β‐Ala‐Pda, 1 (C23H46N2O3: molecular weight = 398.62) have been described. The peptide crystallized in the crystal system triclinic with space group P21: a = 5.116(3) Å, b = 5.6770(10) Å, c = 21.744(5) Å; α = 87.45°, β = 86.87°, γ = 90.0°; Z = 1. An attractive feature of the crystal molecular structure of 1 is the induction of a reasonably extended backbone conformation of the β‐Ala moiety, i.e. the torsion angles φ ≈ ?115°, µ ≈ 173° and ψ ≈ 122°, correspond to skew?, trans and skew+ conformation, respectively, by an unbranched hydrophobic alkyl chain, Pda, which prefers an all‐anti orientation (θ1 ≈ ?153°, θ2 ≈ … θ14 ≈ ±178°). The observation is remarkable because, systematic conformational investigations of short linear β‐Ala peptides of the type Boc‐β‐Ala‐Xaa‐OCH3 (Xaa = Aib or Acc6) have shown that the chemical and stereochemical characters of the neighboring moieties may be critical in dictating the overall folded and/or unfolded conformational features of the β‐Ala residue. The overall conformation of 1 is typical of a ‘bar’. It appears convincing that, in addition to a number of hydrophobic contacts between the parallel arranged molecules, an array of conventional N‐H…O=C intermolecular H‐bonding interactions stabilize the crystal molecular structure. Moreover, the resulting 14‐membered pseudo‐ring motif, generated by the amide–amide interactions between the adjacent molecules, is completely devoid of nonconventional C?H…O interaction. The potentials of the conformational adaptation of the β‐Ala residue, to influence and stabilize different structural characteristics have been highlighted.  相似文献   

2.
Abstract: A series of three homologous dimethyldiamides Ac‐ΔAla‐NMe2, Ac‐l ‐Ala‐NMe2 and Ac‐dl ‐Ala‐NMe2 has been synthesized and the structures of these amides determined from single‐crystal X‐ray diffraction data. To learn more about the conformational preferences of compounds studied, the fully relaxed (φ–ψ) conformational energy maps in vacuo (AM1) of Ac‐ΔAla‐NMe2 and Ac‐l ‐Ala‐NMe2 were obtained, and the calculated minima reoptimized with the DFT/B3LYP/6–31G** method. The crystal‐state results have been compared with the literature data. Ac‐ΔAla‐NMe2 and other α,β‐dehydroamino acid dimethyldiamides, Ac‐ΔXaa‐NMe2 adopt the conservative conformation of the torsion angles φ, ψ = ~ ?45°, ~130°, which are located in the high‐energy region (region H) of Ramachandran diagram. Ac‐l ‐Ala‐NMe2 and Ac‐dl ‐Ala‐NMe2, as well as other saturated amino acid dimethylamides Ac‐l /dl ‐Xaa‐NMe2, present common peptide structures, and no conformational preferences are observed. Molecular packing of the amides analysed reveals two general hydrogen‐bonded motifs. Dehydro and dl ‐species are paired into centrosymmetric dimers, and l ‐compounds form catemers. However, Ac‐ΔAla‐NMe2 and Ac‐DL ‐Ala‐NMe2 constitute exceptions: their molecules also link into catemers.  相似文献   

3.
Abstract: The structural perturbation induced by CαH→Nα exchange in azaamino acid‐containing peptides was predicted by ab initio calculation of the 6‐31G* and 3‐21G* levels. The global energy‐minimum conformations for model compounds, For‐azaXaa‐NH2 (Xaa = Gly, Ala, Leu) appeared to be the β‐turn motif with a dihedral angle of φ = ± 90°, ψ = 0°. This suggests that incorporation of the azaXaa residue into the i + 2 position of designed peptides could stabilize the β‐turn structure. The model azaLeu‐containing peptide, Boc‐Phe‐azaLeu‐Ala‐OMe, which is predicted to adopt a β‐turn conformation was designed and synthesized in order to experimentally elucidate the role of the azaamino acid residue. Its structural preference in organic solvents was investigated using 1H NMR, molecular modelling and IR spectroscopy. The temperature coefficients of amide protons, the characteristic NOE patterns, the restrained molecular dynamics simulation and IR spectroscopy defined the dihedral angles [ (φi+1, ψi+1) (φi+2, ψi+2)] of the Phe‐azaLeu fragment in the model peptide, Boc‐Phe‐azaLeu‐Ala‐OMe, as [(?59°, 127°) (107°, ?4°)]. This solution conformation supports a βII‐turn structural preference in azaLeu‐containing peptides as predicted by the quantum chemical calculation. Therefore, intercalation of the azaamino acid residue into the i + 2 position in synthetic peptides is expected to provide a stable β‐turn formation, and this could be utilized in the design of new peptidomimetics adopting a β‐turn scaffold.  相似文献   

4.
[14C]‐N‐(6‐Chloro‐7‐methoxy‐9H‐pyrido [3,4‐b]indol‐8‐yl)‐2‐methyl‐3‐pyridinecarboxamide (5B ), an IKK inhibitor, was synthesized from [14C]‐barium carbonate in two steps in an overall radiochemical yield of 41%. The intermediate, [carboxyl‐14C]‐2‐methylnicotinic acid, was prepared by the lithiation and carbonation of 3‐bromo‐2‐methylpyridine. [13C4,D3]‐N‐(6‐chloro‐7‐methoxy‐9H‐pyrido [3,4‐b]indol‐8‐yl)‐2‐methyl‐3‐pyridinecarboxamide (5C ) was synthesized from [1,2,3,4‐13C4]‐ethyl acetoacetate and [D4]‐methanol in six steps in an overall yield of 2%. [13C4]‐2‐methylnicotic acid, was prepared by condensation of [13C4]‐ethyl 3‐aminocrotonate and acrolein, followed by hydrolysis with lithium hydroxide. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

5.
Abstract: During the manufacture of a proprietary peptide drug substance a new impurity appeared unexpectedly. Investigation of its chemical structure established the impurity as a β‐Ala insertion mutant of the mother peptide. The source of the β‐Ala was identified as contamination of the Fmoc‐Ala‐OH raw material with Fmoc‐β‐Ala‐Ala‐OH. Further studies also demonstrated the presence of β‐Ala in other Fmoc‐amino acids, particularly in Fmoc‐Arg(Pbf)‐OH. In this case, it was due to the presence of both Fmoc‐β‐Ala‐OH and Fmoc‐β‐Ala‐Arg(Pbf)‐OH. It is concluded that β‐Ala contamination of Fmoc‐amino acid derivatives is a general and hitherto unrecognized problem to suppliers of Fmoc‐amino acid derivatives. The β‐Ala is often present as Fmoc‐β‐Ala‐OH and/or as a dipeptide, Fmoc‐β‐Ala‐amino acid‐OH. In collaboration with the suppliers, new specifications were introduced, recognizing the presence of β‐Ala‐related impurities in the raw materials and limiting them to acceptable levels. The implementation of these measures has essentially eliminated β‐Ala contamination as a problem in the manufacture of the drug substance.  相似文献   

6.
Mimetics of β‐turn structures in proteins have been used to calibrate the relative reactivities toward deamidation of asparagine residues in the two central positions of a β‐turn and in a random coil. N‐Acetyl‐Asn‐Gly‐6‐aminocaproic acid, an acyclic analog of a β‐turn mimic undergoes deamidation of the asparaginyl residue through a succinimide intermediate to generate N‐acetyl‐Asp‐N‐Gly‐6‐aminocaproic acid (6‐aminocaproic acid, hereafter Aca) and N‐acetyl‐l ‐iso‐aspartyl (isoAsp)‐Gly‐Aca (pH 8.8, 37 °C) ≈ 3‐fold faster than does the cyclic β‐turn mimic cyclo‐[L‐Asn‐Gly‐Aca] with asparagine at position 2 of the β‐turn. The latter compound, in turn, undergoes deamidation ≈ 30‐fold faster than its positional isomer cyclo‐[Gly‐Asn‐Aca] with asparagine at position 3 of the β‐turn. Both cyclic peptides assume predominantly β‐turn structures in solution, as demonstrated by NMR and circular dichroism characterization. The open‐chain compound and its isomer N‐acetyl‐Gly‐Asn‐Aca assume predominantly random coil structures. The latter isomer undergoes deamidation 2‐fold slower than the former. Thus the order of reactivity toward deamidation is: asparagine in a random coil ≈ 3× asparagine in position 2 of a β‐turn ≈ 30× asparagine in position 3 of a β‐turn.  相似文献   

7.
Abstract: Replacement of Phe3 in the endogenous δ‐opioid selective peptide deltorphin I with four optically pure stereoisomers of the topographically constrained, highly hydrophobic novel amino acid β‐isopropylphenylalanine (β‐iPrPhe) produced four pharmacologically different deltorphin I peptidomimetics. Radiolabeled ligand‐binding assays and in vitro biological evaluation indicate that the stereoconfiguration of the iPrPhe residue plays a crucial role in determining the binding affinity, bioactivity and selectivity of [β‐iPrPhe3]deltorphin I analogs: a (2S,3R) configuration of the iPrPhe3 residue in [β‐iPrPhe3]deltorphin I provided the most desirable biological properties with binding affinity (IC50 = 2 n m ), bioassay potency (IC50 = 1.23 n m in MVD assay) and exceptional selectivity for the δ‐opioid receptor over the µ‐opioid receptor (30 000). Further conformational studies based on two‐dimensional NMR and computer‐assisted molecular modeling suggested a model for the possible bioactive conformation in which the Tyr1 and (2S,3R)‐β‐iPrPhe3 residues adopt trans side‐chain conformations, and the linear peptide backbone favors a distorted β‐turn conformation.  相似文献   

8.
Synthesis of [15N4] purine labeled cytokinine glycosides derived from zeatins and topolins containing a 9‐β‐d , 7‐β‐d ‐glucopyranosyl, or 9‐β‐d ‐ribofuranosyl group is described. These N6‐substituted adenine derivatives are intended as internal analytic standards for phytohormone analysis. All labeled compounds were prepared from 6‐chloro[15N4]purine ( 1 ). The equilibrium reaction of 1 with acetobromo‐α‐d ‐glucose gave isomeric 7‐β‐d ( 3 ) and 9‐β‐d ( 4 ) chloro glucosyl precursors, which were treated with the corresponding amines to get desired labeled cytokinin 7‐β‐d ( 6 ) and 9‐β‐d ( 5 ) glucopyranosides. Cytokinins containing 9‐β‐d ‐ribofuranosyl group ( 8 ) were obtained by direct enzymatic transglycosylation reaction of cytokinins ( 7 ) prepared from 6‐chloro[15N4] purine ( 1 ).  相似文献   

9.
SSR180575 (7‐chloro‐N,N,5‐trimethyl‐4‐oxo‐3‐phenyl‐3,5‐dihydro‐4H‐pyridazino[4,5‐b]indole‐1‐acetamide) is the lead compound of an original pyridazinoindole series of potent and highly selective TSPO (peripheral benzodiazepine receptor) ligands. Isotopic labeling of SSR180575 with the short‐lived positron‐emitter carbon‐11 (T1/2: 20.38 min) at its 5‐methylpyridazino[4,5‐b]indole moiety as well as at its N,N‐dimethylacetamide function by methylation of the corresponding nor‐analogues was investigated. Best results in terms of radiochemical yields and purities were obtained for the preparation of [indole‐N‐methyl‐11C]SSR180575, where routine production batches of 4.5–5.0 GBq of radiochemically pure (>99%) i.v. injectable solutions (specific radioactivities: 50–90 GBq/ µ mol) could be prepared within a total synthesis time of 25 min (HPLC purification included) starting from a 55 GBq [11C]CO2 cyclotron production batch (non‐decay‐corrected overall radiochemical yields: 8–9%). The process comprises (1) trapping at ?10°C of [11C]methyl triflate in DMF (300 µ l) containing 0.2–0.3 mg of the indole precursor for labeling and 4 mg of K2CO3 (excess); (2) heating at 120°C for 3 min; (3) dilution of the residue with 0.5 ml of the HPLC mobile phase and (4) purification using semi‐preparative reversed‐phase HPLC (Zorbax® SB‐C‐18). In vivo pharmacological properties of [indole‐N‐methyl‐11C]SSR180575 as a candidate for imaging neuroinflammation with positron emission tomography are currently evaluated. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

10.
Abstract: The structure of the peptide Boc‐Ala‐Leu‐Ac7c‐Ala‐Leu‐Ac7c‐OMe (Ac7c,1‐aminocycloheptane‐1‐carboxylic acid) is described in crystals. The presence of two Ac7c residues was expected to stabilize a 310‐helical fold. Contrary to expectation the structural analysis revealed an unfolded amino terminus, with Ala(1) adopting an extended β‐conformation (φ = ?93°,ψ = 112°). Residues 2–5 form a 310‐helix, stabilized by three successive intramolecular hydrogen bonds. Notably, two NH groups Ala(1) and Ac7c(3) do not form any hydrogen bonds in the crystal. Peptide assembly appears to be dominated by packing of the cycloheptane rings that stack against one another within the molecule and also throughout the crystal in columns.  相似文献   

11.
Abstract: Two complete series of N‐protected oligopeptide esters to the pentamer level from 1‐amino‐cyclodecane‐1‐carboxylic acid (Ac10c), an α‐amino acid conformationally constrained through a medium‐ring Cαi ? Cαi cyclization, and either the l ‐Ala or Aib residue, along with the N‐protected Ac10c monomer and homo‐dimer alkylamides, were synthesized using solution methods and fully characterized. The preferred conformation of these model peptides was assessed in deuterochloroform solution using FT‐IR absorption and 1H NMR techniques. Furthermore, the molecular structures of two derivatives (Z‐Ac10c‐OH and Fmoc‐Ac10c‐OH) and two peptides (the dipeptide ester Z‐Ac10c‐l ‐Phe‐OMe and the tripeptide ester Z‐Aib‐Ac10c‐Aib‐OtBu) were determined in the crystal state using X‐ray diffraction. The experimental results support the view that β‐bends and 310‐helices are preferentially adopted by peptides rich in Ac10c, the third largest cycloaliphatic Cα,α‐disubstituted glycine known. This investigation allowed us to complete a detailed conformational analysis of the whole 1‐amino‐cycloalkane‐1‐carboxylic acid (Acnc, with n = 3–12) series, which represents the prerequisite for our recent proposal of the ‘Acnc scan’ concept.  相似文献   

12.
Abstract: The design, synthesis, characterization and self‐assembling properties of a new class of amphiphilic peptides, constructed from a bifunctional polar core attached to totally hydrophobic arms, are presented. The first series of this class, represented by the general structure Py(Aibn)2 (Py = 2,6‐pyridine dicarbonyl unit; Aib = α, α′‐dimethyl glycine; n = 1–4), is prepared in a single step by the condensation of commercially available 2,6‐pyridine dicarbonyl dichloride with the methyl ester of homo oligoAib peptide (Aibn‐OMe) in the presence of triethyl amine. 1H NMR VT and ROESY studies indicated the presence of a common structural feature of 2‐fold symmetry and an NH…N hydrogen bond for all the members. Whereas the Aib3 segment in Py(Aib3)2 showed only the onset of a 310‐helical structure, the presence of a well‐formed 310‐helix in both Aib4 arms of Py(Aib4)2 was evident in the 1H NMR of the bispeptide. X‐ray crystallographic studies have shown that in the solid state, whereas Py(Aib2)2 molecules organize into a sheet‐like structure and Py(Aib3)2 molecules form a double‐stranded string assembly, the tetra Aib bispeptide, Py(Aib4)2, is organized to form a tetrameric assembly which in turn extends into a continuous channel‐like structure. The channel is totally hydrophobic in the interior and can selectively encapsulate lipophilic ester (CH3COOR, R = C2H5, C5H11) molecules, as shown by the crystal structures of the encapsulating channel. The crystal structure parameters are: 1b , Py(Aib2)2, C25H37N5O8, sp. gr. P212121, a = 9.170(1) Å, b = 16.215(2) Å, c = 20.091(3) Å, R = 4.80; 1c , Py(Aib3)2, C33H51N7O10·H2O, sp. gr. P, a = 11.040(1) Å, b = 12.367(1) Å, c = 16.959(1) Å, α = 102.41°, β = 97.29°, γ = 110.83°, R1 = 6.94; 1 da, Py(Aib4)2?et ac, C41H65N9O12?1.5H2O·C4H8O2, sp. gr. P, a = 16.064(4) Å, b = 16.156 Å, c = 21.655(5) Å, α = 90.14(1)°, β = 101.38(2)°, γ = 97.07(1)°, Z = 4, R1 = 9.03; 1db, Py(Aib4)2?amylac,C41H65N9O12?H2O ·C7H14O2, P21/c, a = 16.890(1) Å, b = 17.523(1) Å, c = 20.411(1) Å, β = 98.18 °, Z = 4, R = 11.1 (with disorder).  相似文献   

13.
Abstract: The synthesis of conformationally restricted dipeptidic moieties 4‐amino‐1,2,4,5‐tetrahydro‐2‐benzazepin‐3‐one (Aba)‐Gly ([(4S)‐amino‐3‐oxo‐1,2,4,5‐tetrahydro‐1H‐2‐benzazepin‐2‐yl]‐acetic acid) and 8‐hydroxy‐4‐amino‐1,2,4,5‐tetrahydro‐2‐benzazepin‐3‐one (Hba)‐d ‐Ala ([(4S)‐amino‐8‐hydroxy‐3‐oxo‐1,2,4,5‐tetrahydro‐benzo[c]azepin‐2‐yl]‐propionic acid) was based on a synthetic strategy that uses an oxazolidinone as an N‐acyliminium precursor. Introducing these Aba scaffolds into the N‐terminal tetrapeptide of dermorphin (H‐Tyr‐d ‐Ala‐Phe‐Gly‐Tyr‐Pro‐Ser‐NH2)‐induced remarkable shifts in affinity and selectivity towards the opioid μ‐ and δ‐receptors. This paper provides the synthesis and biological in vitro and in vivo evaluation of constricted analogues of the N‐terminal tetrapeptide H‐Tyr‐d ‐Ala‐Phe‐Gly‐NH2, which is the minimal subunit of dermorphin needed for dermorphin‐like opiate activity.  相似文献   

14.
14C‐Labelled N‐(2‐chloro‐3,4‐dimethoxybenzylideneamino)guanidinium acetate has been synthesized as a part of a four‐step procedure which involved decarboxylation of 2‐chloro‐3,4‐dimethoxybenzoic acid by Pb(OAc)4 to give 2‐chloro‐3,4‐dimethoxy‐1‐iodobenzene, followed by a selective lithiation at the iodine position and electrophilic substitution with N,N‐dimethylformamide [α14C] and final reaction with aminoguanidine bicarbonate. The specific activity was 59 mCi/mmol and the overall yield 49%. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

15.
Abstract: In order to make clear the structural role of the C‐terminal amide group of endomorphin‐2 (EM2, H‐Tyr‐Pro‐Phe‐Phe‐NH2), an endogenous µ‐receptor ligand, in the biological function, the solution conformations of endomorphin‐2 and its C‐terminal free acid (EM2OH, H‐Tyr‐Pro‐Phe‐Phe‐OH), studied using two‐dimensional 1H NMR measurements and molecular modeling calculations, were compared. Both peptides were in equilibrium between the cis and trans isomers around the Tyr‐Pro ω bond in a population ratio of ≈ 1 : 2. The lack of significant temperature and concentration dependence of NH protons suggested that the NMR spectra reflected the conformational features of the respective molecules themselves. Fifty possible 3D structures for the each isomer were generated by the dynamical simulated annealing method under the proton?proton distance constraints derived from the ROE cross‐peaks. These energy‐minimized conformers, which were all in the φ torsion angles estimated from JNHCαH coupling constants within ± 30°, were then classified in groups one or two according to the folding backbone structures. All trans and cis EM2 conformers adopt an open conformation in which their extended backbone structures are twisted at the Pro2–Phe3 moiety. In contrast, the trans and cis conformers of EM2OH show conformational variation between the ‘bow’‐shaped extended and folded backbone structures, although the cis conformers of its zwitterionic form are refined into the folded structure of the close disposition of C‐ and N‐terminal groups. These results indicate clearly that the substitution of carboxyl group for C‐terminal amide group makes the peptide flexible. The conformational requirement for µ‐receptor activation has been discussed based on the active form proposed for endomorphin‐1 and by comparing conformational features of EM2 and EM2OH.  相似文献   

16.
The dehydro-peptide Boc-L-Ile-ΔPhe-L-Trp-OCH3 was synthesized by the azlactone method in the solution phase. The peptide was crystallized from methanol in an orthorhombic space group P212121 with a = 10.777(2), b= 11.224(2), c= 26.627(10) Å. The structure was determined by direct methods and refined to an R value of 0.069 for 3093 observed reflections [l≥ 2σ(l)].The peptide failed to adopt a folded conformation with backbone torsion angles: φ1, = 90.8(8)°, ψ1= -151.6(6)°, φ2= 89.0(8)°, ψ2= 15.9(9)°, φ3= 165.7(7)°, ψT3= -166.0(7)°. A general rule derived from earlier studies indicates that a three-peptide unit sequence with a ΔPhe at the (i+ 2) position adopts a β-turn II conformation. Because the branched β-carbon residues such as valine and isoleucine have strong conformational preferences, they combine with the ΔPhe residue differently to generate a unique set of conformations in such peptides. The presence of β-branched residues simultaneously at both (i+ 1) and (i+ 3) positions induces unfolded conformations in tetrapeptides, but a β-branched residue substituted only at (i+ 3) positron can not prevent the formation of a folded β-turn II conformation. On the other hand, the present structure shows that a β-branched residue substituted at the (i+ 1) position prevents the formation of a β-turn II conformation. These observations indicate that a β-branched residue at the (i+ 1) position prevents a folded conformation whereas it cannot generate the same degree of effect from the (i+ 3) position. This may be because of the trans disposition of the planar ΔPhe side-chain with respect to the C=O group in the residue. The molecules are packed in an anti-parallel manner to generate N2-H2…O2 (-x,y-1/2, -z+ 3/2) and Nε13-Hε13…O1(-x,y -1/2, -z+ 3/2) hydrogen bonds.  相似文献   

17.
Abstract: A spontaneously folding β‐hairpin peptide (Lys‐Lys‐Tyr‐Thr‐Val‐Ser‐Ile‐Asn‐Gly‐Lys‐Lys‐Ile‐Thr‐Val‐Ser‐Ile) and related cyclic (cyclo‐Gly‐Lys‐Tyr‐Ile‐Asn‐Gly‐Lys‐Ile‐Ile‐Asn) and linear (Ser‐Ile‐Asn‐Gly‐Lys) controls were studied to determine the effects of various factors on secondary structure. Secondary structure was evaluated using circular dichroism (CD) and 1D and 2D 1H nuclear magnetic resonance (NMR). The effects of chemical modifications in the peptide and various solution conditions were investigated to determine their impact on peptide structure. The β‐hairpin peptide displayed a CD minimum at 216 nm and a TOCSY i + 1 ? i + 2 and i + 2 ?i + 3 interaction, confirming the expected structure. Using NMR α‐proton (H) chemical shifts, the extents of folding of the β‐hairpin and linear control were estimated to be 51 and 25% of the cyclic control (pH 4, 37 °C), which was taken to be maximally folded. Substitution of iso‐aspartic acid for Asn reduced the secondary structure dramatically; substitution of aspartic acid for Asn also disrupted the structure. This result suggests that deamidation in unconstrained β‐turns may have adverse effects on secondary structure. N‐terminal acetylation and extreme pH conditions also reduced structure, while the addition of methanol increased structure.  相似文献   

18.
New derivatives with the tetrahydro‐β‐carboline‐imidazolidinedione and tetrahydro‐β‐carboline‐piperazinedione scaffolds and a pendant bromothienyl moiety at C‐5/C‐6 were synthesized and tested for their ability to inhibit PDE5 in vitro. The following SAR can be concluded: The tetracyclic scaffold is essential for PDE5 inhibition; the ethyl group is the most suitable among the adopted N‐substituents on the terminal ring (hydantoin/piperazinedione); the appropriate stereochemistry of C‐5/C‐6 derived from the aldehyde rather than C‐11a/C‐12a derived from tryptophan appears crucial for inhibition of PDE5; surprisingly, derivatives with the hydantoin terminal ring are more active than their analogs with the piperazinedione ring; the selectivity versus PDE5 relative to PDE11 with cGMP as a substrate is mainly a function of the substitution and stereochemistry pattern of the external ring, in other words of the interaction with the H‐loop residues of the isozymes. Thirteen derivatives showed PDE5 inhibitory activity with IC50 values in the range of 0.16–5.4 µm. Compound 8 was the most potent PDE5 inhibitor and showed selectivity towards PDE5 versus other PDEs, with a selectivity index of 49 towards PDE5 rather than PDE11 with cGMP as the substrate.  相似文献   

19.
Abstract: We describe here a systematic study to determine the effect on secondary structure of d ‐amino acid substitutions in the nonpolar face of an amphipathic α‐helical peptide. The helix‐destabilizing ability of 19 d ‐amino acid residues in an amphipathic α‐helical model peptide was evaluated by reversed‐phase HPLC and CD spectroscopy. l ‐Amino acid and d ‐amino acid residues show a wide range of helix‐destabilizing effects relative to Gly, as evidenced in melting temperatures (ΔTm) ranging from ?8.5°C to 30.5°C for the l ‐amino acids and ?9.5°C to 9.0°C for the d ‐amino acids. Helix stereochemistry stability coefficients defined as the difference in Tm values for the l ‐ and d ‐amino acid substitutions [(ΔTm′ = TmL and TmD)] ranging from 1°C to 34.5°C. HPLC retention times [ΔtR(XL?XD)] also had values ranging from ?0.52 to 7.31 min at pH 7.0. The helix‐destabilizing ability of a specific d ‐amino acid is highly dependent on its side‐chain, with no clear relationship to the helical propensity of its corresponding l ‐enantiomers. In both CD and reversed‐phase HPLC studies, d ‐amino acids with β‐branched side‐chains destabilize α‐helical structure to the greatest extent. A series of helix stability coefficients was subsequently determined, which should prove valuable both for protein structure‐activity studies and de novo design of novel biologically active peptides.  相似文献   

20.
Recently, a novel series of 2‐phenylpyrazolo[1,5‐a]pyrimidineacetamides has been reported as selective ligands of the translocator protein (18 kDa). Within this series, DPA‐714 (N,N‐diethyl‐2‐(2‐(4‐(2‐fluoroethoxy)phenyl)‐5,7‐dimethylpyrazolo[1,5‐a]pyrimidin‐3‐yl)acetamide, Ki=7.0 nM) is a compound, which had been designed with a fluorine atom in its structure, allowing labelling with fluorine‐18 (half‐life: 109.8 min) and in vivo imaging using positron emission tomography. DPA‐714 and its tosyloxy derivative (N,N‐diethyl‐2‐(2‐(4‐(2‐toluenesulfonyloxyethoxy)phenyl)‐5,7‐dimethylpyrazolo[1,5‐a]pyrimidin‐3‐yl)acetamide) as precursor for the labelling with fluorine‐18 were synthesized in two steps from DPA‐713 (N,N‐diethyl‐2‐(2‐(4‐methoxyphenyl)‐5,7‐dimethylpyrazolo[1,5‐a]pyrimidin‐3‐yl)acetamide) and obtained in 32 and 42% yields, respectively. [18F]DPA‐714 was synthesized using a simple one‐step process (a tosyloxy‐for‐fluorine nucleophilic aliphatic substitution), which has been fully automated on our Zymate‐XP robotic system. It involves: (A) reaction of K[18F]F‐Kryptofix®222 with the tosyloxy precursor (4.5–5.0 mg, 8.2–9.1 µmol) at 165°C for 5 min in dimethyl sufloxide (0.6 mL) followed by (B) C18 PrepSep cartridge pre‐purification and finally (C) semi‐preparative high‐performance liquid chromatography (HPLC) purification on a Waters X‐Terra? RP18. Typically, 5.6–7.4 GBq of [18F]DPA‐714 (>95% chemically and radiochemically pure) could be obtained with specific radioactivities ranging from 37 to 111 GBq/µmol within 85–90 min (HPLC purification and SepPak®‐based formulation included), starting from a 37 GBq [18F]fluoride batch (overall non‐decay‐corrected and isolated radiochemical yield: 15–20%). Copyright © 2008 John Wiley & Sons, Ltd.  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