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 210 毫秒
1.
Baysal C  Meirovitch H 《Biopolymers》2000,54(6):416-428
A statistical mechanics methodology for predicting the solution structures and populations of peptides developed recently is based on a novel method for optimizing implicit solvation models, which was applied initially to a cyclic hexapeptide in DMSO (C. Baysal and H. Meirovitch, Journal of American Chemical Society, 1998, vol. 120, pp. 800-812). Thus, the molecule has been described by the simplified energy function E(tot) = E(GRO) + summation operator(k) sigma(k)A(k), where E(GRO) is the GROMOS force-field energy, sigma(k) and A(k) are the atomic solvation parameter (ASP) and the solvent accessible surface area of atom k, respectively. In a more recent study, these ASPs have been found to be transferable to the cyclic pentapeptide cyclo(D-Pro(1)-Ala(2)-Ala(3)-Ala(4)-Ala(5)) in DMSO (C. Baysal and H. Meirovitch, Biopolymers, 2000, vol. 53, pp. 423-433). In the present paper, our methodology is applied to the cyclic heptapeptides axinastatin 2 [cyclo(Asn(1)-Pro(2)-Phe(3)-Val(4)-Leu(5)-Pro(6)-Val(7))] and axinastatin 3 [cyclo(Asn(1)-Pro(2)-Phe(3)-Ile(4)-Leu(5)-Pro(6)-Val(7))], in DMSO, which were studied by nmr by Mechnich et al. (Helvetica Chimica Acta, 1997, vol. 80, pp. 1338-1354). The calculations for axinastatin 2 show that special ASPs should be optimized for the partially charged side-chain atoms of Asn while the rest of the atoms take their values derived in our previous work; this suggests that similar optimization might be needed for other side chains as well. The solution structures of these peptides are obtained ab initio (i.e., without using experimental restraints) by an extensive conformational search based on E(GRO) alone and E(*)(tot), which consists of the new set of ASPs. For E(*)(tot), the theoretical values of proton-proton distances, (3)J coupling constants, and other properties are found to agree very well with the nmr results, and they are always better than those based on E(GRO).  相似文献   

2.
Das B  Meirovitch H 《Proteins》2003,51(3):470-483
A new procedure for optimizing parameters of implicit solvation models introduced by us has been applied successfully first to cyclic peptides and more recently to three surface loops of ribonuclease A (Das and Meirovitch, Proteins 2001;43:303-314) using the simplified model E(tot) = E(FF)(epsilon = nr) + Sigma(i) sigma(i)A(i), where sigma(i) are atomic solvation parameters (ASPs) to be optimized, A(i) is the solvent accessible surface area of atom i, E(FF)(epsilon = nr) is the AMBER force-field energy of the loop-loop and loop-template interactions with a distance-dependent dielectric constant, epsilon = nr, where n is a parameter. The loop is free to move while the protein template is held fixed in its X-ray structure; an extensive conformational search for energy minimized loop structures is carried out with our local torsional deformation method. The optimal ASPs and n are those for which the structure with the lowest minimized energy [E(tot)(n,sigma(i))] becomes the experimental X-ray structure, or less strictly, the energy gap between these structures is within 2-3 kcal/mol. To check if a set of ASPs can be defined, which is transferable to a large number of loops, we optimize individual sets of ASPs (based on n = 2) for 12 surface loops from which an "averaged" best-fit set is defined. This set is then applied to the 12 loops and an independent "test" group of 8 loops leading in most cases to very small RMSD values; thus, this set can be useful for structure prediction of loops in homology modeling. For three loops we also calculate the free energy gaps to find that they are only slightly smaller than their energy counterparts, indicating that only larger n will enable reducing too large gaps. Because of its simplicity, this model allowed carrying out an extensive application of our methodology, providing thereby a large number of benchmark results for comparison with future calculations based on n > 2 as well as on more sophisticated solvation models with as yet unknown performance for loops.  相似文献   

3.
Das B  Meirovitch H 《Proteins》2001,43(3):303-314
A novel procedure for optimizing the atomic solvation parameters (ASPs) sigma(i) developed recently for cyclic peptides is extended to surface loops in proteins. The loop is free to move, whereas the protein template is held fixed in its X-ray structure. The energy is E(tot) = E(FF)(epsilon = nr) + summation operator sigma(i)A(i), where E(FF)(epsilon = nr) is the force-field energy of the loop-loop and loop-template interactions, epsilon = nr is a distance-dependent dielectric constant, and n is an additional parameter to be optimized. A(i) is the solvent-accessible surface area of atom i. The optimal sigma(i) and n are those for which the loop structure with the global minimum of E(tot)(n, sigma(i)) becomes the experimental X-ray structure. Thus, the ASPs depend on the force field and are optimized in the protein environment, unlike commonly used ASPs such as those of Wesson and Eisenberg (Protein Sci 1992;1:227-235). The latter are based on the free energy of transfer of small molecules from the gas phase to water and have been traditionally combined with various force fields without further calibration. We found that for loops the all-atom AMBER force field performed better than OPLS and CHARMM22. Two sets of ASPs [based on AMBER (n = 2)], optimized independently for loops 64-71 and 89-97 of ribonuclease A, were similar and thus enabled the definition of a best-fit set. All these ASPs were negative (hydrophilic), including those for carbon. Very good (i.e., small) root-mean-square-deviation values from the X-ray loop structure were obtained with the three sets of ASPs, suggesting that the best-fit set would be transferable to loops in other proteins as well. The structure of loop 13-24 is relatively stretched and was insensitive to the effect of the ASPs.  相似文献   

4.
5.
Base stacking is a complicated and not completely understood phenomenon that is influenced by contributions of electrostatic (dipole-dipole and dipole-induced dipole) interactions, dispersion (permanent dipole-induced dipole) effects and solvation effects. The plots of those factors did not show qualitative correlation (Guckian et al., J. Am. Chem. Soc. 1996, 118, 8182-8183). We tried to correlate the stacking and solvation contributions with lipophilicity, extent of fluorine substitution and dipole moment.  相似文献   

6.
Modifications to the AMBER force field [W.D. Cornell, P. Cieplak, C.I. Bayly, I.R. Gould, K. Merz, D.M. Ferguson, D.C. Spellmeyer, T. Fox, J.W. Caldwell, P.A. Kollman, J. Am. Chem. Soc., 117 (1995) 5179-5197] have been made to improve our ability to reproduce observed molecular properties of alpha-linked carbohydrates when calculated using empirical potential-energy functions. Molecular structures and energies obtained using gradient-optimized density functional methods with ab initio basis sets (B3LYP/6-31G*) on ten minimum-energy conformations of maltose [F.A. Momany, J.L. Willett, J. Comp. Chem., submitted for publication] were used to refine the empirical potentials. Molecular dynamics simulations on beta-maltose (i.e., the beta anomer of maltose), cyclohexamylose (alpha-cyclodextrin), cycloheptamylose (beta-cyclodextrin) and larger cyclomaltooligosaccharide structures were carried out and compared with experimental structural studies to test the new potentials. Ring-puckering potential during dynamics as well as conformational transitions to 'flipped' structures were examined. Results of the tests described here suggest that the revised AMBER parameters (AMB99C) are very good for computational studies of alpha-(1-->4)-linked carbohydrates.  相似文献   

7.
From calculations of a model reaction scheme for base-catalyzed RNA hydrolysis, a pentacoodinate dianionic intermediate 2a (Storer, et al., J. Am. Chem. Soc., 1991, 113, 5216-5219) as well as two transition states, TS1 and TS2, to the intermediate have been located by ab initio calculations at the 3-21G* level. Although the intermediate, which has the well depth on the order of kBT, is unlikely to be kinetically significant, the overall rate-limiting transition state structure TS2 obtained at 3-21G* level is very close to the corresponding structure at the STO-3G level; it has an extended P-O(5') bond breaking character. These gas-phase calculation results are used to qualitatively interpret mutagenesis results of Barnase and RNase T1 where water molecules are absent from the active site.  相似文献   

8.
van Dijk AD  Fushman D  Bonvin AM 《Proteins》2005,60(3):367-381
When classical, Nuclear Overhauser Effect (NOE)-based approaches fail, it is possible, given high-resolution structures of the free molecules, to model the structure of a complex in solution based solely on chemical shift perturbation (CSP) data in combination with orientational restraints from residual dipolar couplings (RDCs) when available. RDCs can be incorporated into the docking following various strategies: as direct restraints and/or as intermolecular intervector projection angle restraints (Meiler et al., J Biomol NMR 2000;16:245-252). The advantage of the latter for docking is that they directly define the relative orientation of the molecules. A combined protocol in which RDCs are first introduced as intervector projection angle restraints and at a later stage as direct restraints is shown here to give the best performance. This approach, implemented in our information-driven docking approach HADDOCK (Dominguez et al., J Am Chem Soc 2003;125:1731-1737), is used to determine the solution structure of the Lys48-linked di-ubiquitin, for which chemical shift mapping, RDCs, and (15)N-relaxation data have been previously obtained (Varadan et al., J Mol Biol 2002;324:637-647). The resulting structures, derived from CSP and RDC data, are cross-validated using (15)N-relaxation data. The solution structure differs from the crystal structure by a 20 degrees rotation of the two ubiquitin units relative to each other.  相似文献   

9.
This article concerns the study of the folding-unfolding (Continuous Symmetry Measure, CSM) method of Avnir et al. (Zabrodsky, H.; Peleg, S.; Avnir, D. J. Am. Chem. Soc. 1992, 114, 7843) for quantitative evaluation of the asymmetry of molecular objects. It is shown that a series of modifications to the folding-unfolding method are expedient. An efficient solution for optimisation problems in the folding-unfolding method has been proposed. It allows a significant speed up of the calculations and provides better results. Differences in the behaviours of the original and optimised folding-unfolding methods for various molecular structures are investigated.  相似文献   

10.
Pitera JW  Kollman PA 《Proteins》2000,41(3):385-397
We have extended and applied a multicoordinate free energy method, chemical Monte Carlo/Molecular Dynamics (CMC/MD), to calculate the relative free energies of different amino acid side-chains. CMC/MD allows the calculation of the relative free energies for many chemical species from a single free energy calculation. We have previously shown its utility in host:guest chemistry (Pitera and Kollman, J Am Chem Soc 1998;120:7557-7567)1 and ligand design (Eriksson et al., J Med Chem 1999;42:868-881)2, and here demonstrate its utility in calculations of amino acid properties and protein stability. We first study the relative solvation free energies of N-methylated and acetylated alanine, valine, and serine amino acids. With careful inclusion of rotameric states, internal energies, and both the solution and vacuum states of the calculation, we calculate relative solvation free energies in good agreement with thermodynamic integration (TI) calculations. Interestingly, we find that a significant amount of the unfavorable solvation of valine seen in prior work (Sun et al., J Am Chem Soc 1992;114:6798-6801)3 is caused by restraining the backbone in an extended conformation. In contrast, the solvation free energy of serine is calculated to be less favorable than expected from experiment, due to the formation of a favorable intramolecular hydrogen bond in the vacuum state. These monomer calculations emphasize the need to accurately consider all significant conformations of flexible molecules in free energy calculations. This development of the CMC/MD method paves the way for computations of protein stability analogous to the biochemical technique of "exhaustive mutagenesis." We have carried out just such a calculation at position 133 of T4 lysozyme, where we use CMC/MD to calculate the relative stability of eight different side-chain mutants in a single free energy calculation. Our T4 calculations show good agreement with the prior free energy calculations of Veenstra et al. (Prot Eng 1997;10:789-807)4 and excellent agreement with the experiments of Mendel et al. (Science 1992;256:1798-1802).  相似文献   

11.
Recent studies of noncrystalline HIV-1 capsid protein (CA) assemblies by our laboratory and by Polenova and coworkers (Protein Sci 19:716–730, 2010; J Mol Biol 426:1109–1127, 2014; J Biol Chem 291:13098–13112, 2016; J Am Chem Soc 138:8538–8546, 2016; J Am Chem Soc 138:12029–12032, 2016; J Am Chem Soc 134:6455–6466, 2012; J Am Chem Soc 132:1976–1987, 2010; J Am Chem Soc 135:17793–17803, 2013; Proc Natl Acad Sci USA 112:14617–14622, 2015; J Am Chem Soc 138:14066–14075, 2016) have established the capability of solid state nuclear magnetic resonance (NMR) measurements to provide site-specific structural and dynamical information that is not available from other types of measurements. Nonetheless, the relatively high molecular weight of HIV-1 CA leads to congestion of solid state NMR spectra of fully isotopically labeled assemblies that has been an impediment to further progress. Here we describe an efficient protocol for production of segmentally labeled HIV-1 CA samples in which either the N-terminal domain (NTD) or the C-terminal domain (CTD) is uniformly 15N,13C-labeled. Segmental labeling is achieved by trans-splicing, using the DnaE split intein. Comparisons of two-dimensional solid state NMR spectra of fully labeled and segmentally labeled tubular CA assemblies show substantial improvements in spectral resolution. The molecular structure of HIV-1 assemblies is not significantly perturbed by the single Ser-to-Cys substitution that we introduce between NTD and CTD segments, as required for trans-splicing.  相似文献   

12.
J R Miller  D E Edmondson 《Biochemistry》1999,38(41):13670-13683
Monoamine oxidase A (MAO A) plays a central role in the oxidation of amine neurotransmitters. To investigate the structure and mechanism of this enzyme, recombinant human liver MAO A was expressed and purified from Saccharomyces cerevisiae. Anaerobic titrations of the enzyme require only 1 mol of substrate per mole of enzyme-bound flavin for complete reduction. This demonstrates that only one redox-active group (i.e., the covalent FAD cofactor) is involved in catalysis. The reaction rates and binding affinities of 17 para-substituted benzylamine analogues with purified MAO A were determined by steady state and stopped flow kinetic experiments. For each substrate analogue that was tested, the rates of steady state turnover (k(cat)) and anaerobic flavin reduction (k(red)) are similar in value. Deuterium kinetic isotope effects on k(cat), k(red), k(cat)/K(m), and k(red)/K(s) with alpha, alpha-[(2)H]benzylamines are similar for each substrate analogue that was tested and range in value from 6 to 13, indicating that alpha-C-H bond cleavage is rate-limiting in catalysis. Substrate analogue dissociation constants determined from reductive half-reaction experiments as well as from steady state kinetic isotope effect data [Klinman, J. P., and Matthews, R. G. (1985) J. Am. Chem. Soc. 107, 1058-1060] are in excellent agreement. Quantitative structure-activity relationship (QSAR) analysis of dissociation constants shows that the binding of para-substituted benzylamine analogues to MAO A is best correlated with the van der Waals volume of the substituent, with larger substituents binding most tightly. The rate of para-substituted benzylamine analogue oxidation and/or substrate analogue-dependent flavin reduction is best correlated with substituent electronic effects (sigma). Separation of the electronic substituent parameter (sigma) into field-inductive and resonance effects provides a more comprehensive treatment of the electronic correlations. The positive correlation of rate with sigma (rho approximately 2.0) suggests negative charge development at the benzyl carbon position occurs and supports proton abstraction as the mode of alpha-C-H bond cleavage. These results are discussed in terms of several mechanisms proposed for MAO catalysis and with previous structure-activity studies published with bovine liver MAO B [Walker, M. C., and Edmondson, D. E. (1994) Biochemistry 33, 7088-7098].  相似文献   

13.
Feng Y  Hofstein HA  Zwahlen J  Tonge PJ 《Biochemistry》2002,41(42):12883-12890
Enoyl-CoA hydratase catalyzes the hydration of trans-2-crotonyl-CoA to 3(S)-HB-CoA, 3(S)-hydroxybutyryl-CoA with a stereospecificity (k(S)/k(R)) of 400000 to 1 [Wu, W. J., Feng, Y., He, X., Hofstein, H. S., Raleigh, D. P., and Tonge, P. J. (2000) J. Am. Chem. Soc. 122, 3987-3994]. Replacement of E164, one of the catalytic glutamates in the active site, with either aspartate or glutamine reduces the rate of formation of the 3(S) product enantiomer (k(S)) without affecting the rate of formation of the 3(R) product (k(R)). Consequently, k(S)/k(R) is 1000 and 0.33 for E164D and E164Q, respectively. In contrast, mutagenesis of E144, the second catalytic glutamate, reduces the rate of formation of both product enantiomers. Thus, only E144 is required for the formation of 3(R)-HB-CoA, 3(R)-hydroxybutyryl-CoA. Modeling studies together with analysis of alpha-proton exchange rates and experiments with crotonyl-oxyCoA, a substrate analogue in which the alpha-proton acidity has been reduced 10000-fold, support a mechanism of 3(R)-hydroxybutyryl-CoA formation that involves the E144-catalyzed stepwise addition of water to crotonyl-CoA which is bound in an s-trans conformation in the active site. Finally, we also demonstrate that hydrogen bonds in the oxyanion hole, provided by the backbone amide groups of G141 and A98, are important for the formation of both product enantiomers.  相似文献   

14.
A series of non-natural isosteric analogs of the cationic, ion-pairing, natural amino acids arginine and lysine have been synthesized, characterized with regard to relevant physical parameters, and protected for routine inclusion in Merrifield solid-phase synthesis. The design of these molecules is based on the concept of steric inhibition of solvation, in that judicious placement of alkyl groups can destabilize aqueous ion solvation and favor ion-pairing [see Beeson & Dix (1993) J. Am. Chem. Soc. 115, 10275]. When the residues are substituted for the natural amino acids in biologically active peptides, enhanced ion-pairing of the peptides to their receptors to increase the peptides' biological activities can result. The increased lipophilicity of the non-natural residues can also improve pharmacokinetic parameters and agonist/antagonist behaviors of peptides. While the synthesis of the L-series is described, the D-isomers were also prepared using identical chemistry.  相似文献   

15.
Three unrestrained stochastic dynamics simulations have been carried out on the RNA hairpin GGAC[UUCG] GUCC, using the AMBER94 force field (Cornell et al., 1995. J. Am. Chem. Soc. 117:5179-5197) in MacroModel 5.5 (Mohamadi et al., 1990. J. Comp. Chem. 11:440-467) and either the GB/SA continuum solvation model (Still et al., 1990. J. Am. Chem. Soc. 112:6127-6129) or a linear distance-dependent dielectric (1/R) treatment. The linear distance-dependent treatment results in severe distortion of the nucleic acid structure, restriction of all hydroxyl dihedrals, and collapse of the counterion atmosphere over the course of a 5-ns simulation. An additional vacuum simulation without counterions shows somewhat improved behavior. In contrast, the two GB/SA simulations (1.149 and 3.060 ns in length) give average structures within 1.2 A of the initial NMR structure and in excellent agreement with results of an earlier explicit solvent simulation (Miller and Kollman, 1997. J. Mol. Biol. 270:436-450). In a 3-ns GB/SA simulation starting with the incorrect UUCG tetraloop structure (Cheong et al., 1990. Nature. 346:680-682), this loop conformation converts to the correct loop geometry (Allain and Varani, 1995. J. Mol. Biol. 250:333-353), suggesting enhanced sampling relative to the previous explicit solvent simulation. Thermodynamic effects of 2'-deoxyribose substitutions of loop nucleotides were experimentally determined and are found to correlate with the fraction of time the ribose 2'-OH is hydrogen bonded and the distribution of the hydroxyl dihedral is observed in the GB/SA simulations. The GB/SA simulations thus appear to faithfully represent structural features of the RNA without the computational expense of explicit solvent.  相似文献   

16.
Previously, we reported that certain tetrasubstituted 1,3,5-triaryl-4-alkyl-pyrazoles bind to the estrogen receptor (ER) with high affinity (Fink, B. E.; Mortenson, D. S.; Stauffer, S. R.; Aron, Z. D.; Katzenellenbogen, J. A. Chem. Biol. 1999, 6, 205-219; Stauffer, S. R.; Katzenellenbogen, J. A. J. Comb/. Chem. 2000, 2. 318 329; Stauffer, S. R.: Coletta, C. J.: Sun, J.; Tedesco, R., Katzenellenbogen, B. S.; Katzenellenbogen, J. A. J. Med. Chem. 2000, submitted). To investigate how cyclic permutation of the two nitrogen atoms of a pyrazole might affect ER binding affinity, we prepared a new pyrazole core isomer, namely a 1,3,4-triaryl-5-alkyl-pyrazole (2), to compare it with our original pyrazole (1). We also prepared several peripherally matched core pyrazole isomer sets to investigate whether the two pyrazole series share a common binding orientation. Our efficient, regioselective synthetic route to these pyrazoles relies on the acylation of a hydrazone anion, followed by cyclization, halogenation, and Suzuki coupling. We found that the ER accommodates 1,3,4-triaryl-pyrazoles of the isomeric series only somewhat less well than the original 1,3,5-triaryl series, and it appears that both series share a common binding mode. This preferred orientation for the 1,3,5-triaryl-4-alkyl-pyrazoles is supported by binding affinity measurements of analogues in which the phenolic hydroxyl groups were systematically removed from each of the three aryl groups, and the orientation is consistent, as well, with molecular modeling studies. These studies provide additional insight into the design of heterocyclic core structures for the development of high affinity ER ligands by combinatorial methods.  相似文献   

17.
Joint composite-rotation adiabatic-sweep isotope filters are derived by combining the composite-rotation [Stuart AC et al. (1999) J Am Chem Soc 121: 5346–5347] and adiabatic-sweep [Zwahlen C et al. (1997) J Am Chem Soc 119:6711–6721; Kupče E, Freeman R (1997) J Magn Reson 127:36–48] approaches. The joint isotope filters have improved broadband filtration performance, even for extreme values of the one-bond 1H–13C scalar coupling constants in proteins and RNA molecules. An average Hamiltonian analysis is used to describe evolution of the heteronuclear scalar coupling interaction during the adiabatic sweeps within the isotope filter sequences. The new isotope filter elements permit improved selective detection of NMR resonance signals originating from 1H spins attached to an unlabeled natural abundance component of a complex in which the other components are labeled with 13C and 15N isotopes.  相似文献   

18.
In a previous report, Nowick and co-workers described beta-strand mimic A, which duplicates the structure and hydrogen-bonding pattern of one edge of a tetrapeptide in a beta-strand conformation (Nowick, J. S.; Pairish, M.; Lee, I. Q.; Holmes, D. L.; Ziller, J. W. J. Am. Chem. Soc. 1997, 119, 5413). Beta-strand mimic A is composed of a 5-amino-2-methoxybenzoic acid unit linked to a 5-hydrazino-2-methoxybenzamide unit by means of an acylhydrazine group. This paper introduces two related beta-strand mimics (B and C) and reports their comparison to beta-strand mimic A. Beta-strand mimic B is composed of a 5-amino-2-methoxybenzoic acid unit linked by a diacylhydrazine group to a fumaramide unit; beta-strand mimic C is composed of a 5-amino-2-methoxybenzoic acid unit linked by a diacylhydrazine group to a peptide. Beta-strand mimics A-C were connected to tripeptide (Phe-Ile-Leu) groups by means of 1,2-diaminoethane diurea turn units to form artificial beta-sheets 1-3. 1H NMR studies, involving ROESY, chemical shift, coupling constant, and variable temperature experiments, reveal that 1-3 adopt hydrogen-bonded antiparallel beta-sheet conformations and establish that all three templates are viable beta-strand mimics.  相似文献   

19.
Protein prenylation involves the addition of either a farnesyl (C15) or geranylgeranyl (C20) isoprenoid moiety onto the C-terminus of many proteins. This natural modification serves to direct a protein to the plasma membrane of the cell. A recently discovered application of prenylated peptides is that they have inherent cell-penetrating ability, and are hence termed cell penetrating prenylated peptides. These peptides are able to efficiently cross the cell membrane in an ATP independent, non-endocytotic manner and it was found that the sequence of the peptide does not affect uptake, so long as the geranylgeranyl group is still present [Wollack, J. W.; Zeliadt, N. A.; Mullen, D. G.; Amundson, G.; Geier, S.; Falkum, S.; Wattenberg, E. V.; Barany, G.; Distefano, M. D. Multifunctional Prenylated Peptides for Live Cell Analysis. J. Am. Chem. Soc.2009, 131, 7293-7303]. The present study investigates the effect of removing the fluorophore from the peptides and investigating the uptake by confocal microscopy and flow cytometry. Our results show that the fluorophore is not necessary for uptake of these peptides. This information is significant because it indicates that the prenyl group is the major determinant in allowing these peptides to enter cells; the hydrophobic fluorophore has little effect. Moreover, these studies demonstrate the utility of the Cu-catalyzed click reaction for monitoring the entry of nonfluorescent peptides into cells.  相似文献   

20.
An efficient method for the synthesis of 5'-O-monomethoxytrityl-2',3'-dideoxy-2'-fluoro-3'-thioarabinothymidine [(5'MMT)araF-T(3'SH), (5)] and its 3'-phosphoramidite derivative (6) suitable for automated incorporation into oligonucleotides, is demonstrated. A key step in the synthesis involves reaction of 5'-O-MMT-2,3'-O-anhydrothymidine (4) (Eleuteri, A.; Reese, C.B.; Song, Q. J. Chem. Soc. Perkin Trans. 1 1996, 2237 pp.) with sodium thioacetate to give (5'-MMT)araF-T(3'SAc) (5) (Elzagheid, M.I.; Mattila, K.; Oivanen, M.; Jones, B.C.N.M.; Cosstick, L?nnberg, H. Eur. J. Org. Chem. 2000, 1987-1991). This nucleoside was then converted to its corresponding phosphoramidite derivative, 6, as described previously ((a) Sun, S.; Yoshida, A.; Piccirilli, J.A. RNA, 1997, 3, 1352-1363; (b) Matulic-Adamic, J.; Beigelman, L. Helvetica Chemica Acta 1999, 82, 2141-2150: (c) Fettes, K.J.; O'Neil, I.; Roberts, S.M.; Cosstick, R. Nucleosides, Nucleotides and Nucl. Acids 2001, 20, 1351-1354).  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