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.
A novel platform for nucleic acid recognition that integrates the alpha-helix secondary structure of peptides with the codified base-pairing capability of nucleic acids is reported. The resulting alpha-helical peptide nucleic acids (alpha PNAs) are composed of a repeating tetrapeptidyl unit, aa(1)-aa(2)-aa(3)-Ser(B), where aa(1) through aa(3) represent generic ancillary amino acids and B = nucleobases linked to Ser via a methylene bridge. Effective syntheses of constituent Fmoc-protected nucleoamino acids (Fmoc-Ser(B)-OH, where B = thymine, cytosine, and uracil) are described along with a protocol for the solid-phase synthesis of 21mer alpha PNAs containing five such nucleobases. By varying the ancillary amino acids, two distinct classes of alpha PNAs were constructed, having a net charge of -1 or +6, respectively, at physiological pH. The modular nature of the alpha PNA platform was illustrated by the synthesis of symmetrical disulfide-bridged alpha PNA dimers containing 10 nucleobases. Hybridization of these alpha PNAs with ssDNA has been examined by thermal denaturation, gel electrophoresis, and circular dichroism (CD) and the data indicated that alpha PNA binds to ssDNA in a cooperative manner with high affinity and sequence specificity. In general, b2 alpha PNAs bind faster and more strongly with ssDNA than do the corresponding b1 alpha PNAs. Parallel alpha PNA-DNA complexes are more stable than their antiparallel counterparts. CD studies also revealed that the hybridization event involves the folding of both species into their helical conformations. Finally, NMR experiments provided conclusive evidence of Watson-Crick base pairing in alpha PNA-ssDNA hybrids.  相似文献   

2.
The solution structure and the dimerization behavior of the lipophilic, highly C(alpha)-methylated model peptide, mBrBz-Iva(1)-Val(2)-Iva(3)-(alphaMe)Val(4)-(alphaMe)Phe(5)-(alphaMe)Val(6)-Iva(7)-NHMe, was studied by NMR spectroscopy and molecular dynamics simulations. The conformational analysis resulted in a right-handed 3(10)/alpha-helical equilibrium fast on the NMR time scale with a slight preference for the alpha-helical conformation. The NOESY spectrum showed intermolecular NOEs due to an aggregation of the heptapeptide. In addition, temperature-dependent diffusion measurements were performed to calculate the hydrodynamic radius. All these findings are consistent with an antiparallel side-by-side dimerization. The structure of the dimeric peptide was calculated with a simulated annealing strategy. The lipophilic dimer is held together by favorable van der Waals interactions in the sense of a bulge fitting into a groove. The flexibility of the helical conformations concerning an alpha/3(10)-helical equilibrium is shown in a 3 ns molecular dynamics simulation of the resulting dimeric structure. Both overall helical structures of each monomer and the antiparallel mode of dimerization are stable. However, transitions were seen of several residues from a 3(10)-helical into an alpha-helical conformation and vice versa. Hence, this peptide represents a good model in which two often-discussed aspects of hierarchical transmembrane protein folding are present: i <-- i + 3 and i <-- i + 4 local H-bonding interactions cause a specific molecular shape which is then recognized as attractive by other surrounding structures.  相似文献   

3.
The method proposed for the evaluation of statistical weights in paper I, and the three-state model [alpha-helical (alpha), extended (epsilon), and other (c) states] formulated in paper II, have been used to develop a procedure to predict the backbone conformations of proteins, based on the concept of the predominant role played by shortrange interactions in determining protein conformation. Conformational probability profiles, in which the probabilities of formation of three consecutive alpha-helical conformations (triad) and of four consecutive extended conformations (tetrad) have been defined relative to their average values over the whole molecule, are calculated for 19 proteins, of which 16 had been used in paper I to evaluate the set of statistical weights of the 20 naturally occurring amino acids. By comparing these conformational probability profiles to experimental x-ray observations, the following results have been obtained: 80% of the alpha-helical regions and 72% of the extended conformational regions have been predicted correctly for the 19 proteins. The percentage of residues predicted correctly is in the range of 53 to 90% for the alpha-helical conformation and in the range of 63 to 88% for the extended conformation for the 19 proteins in the two-state models [alpha-helical (alpha) and other (c) states, and extended (epsilon) and other (c) states]. In the three-state model, the percentage of residues predicted correctly is in the range of 47% to 77 for 19 proteins. These results suggest that the assumption of the dominance of short-range interactions, on which the predictive scheme is based, is a reasonable one. The present predictive method is compared with that of other authors.  相似文献   

4.
A vibrational Raman optical activity (ROA) study, supplemented by protein X-ray crystal structure data, of alpha-helices in polypeptides, proteins, and viruses has suggested that ROA bands in the extended amide III spectral region may be used to distinguish between two types of right-handed alpha-helix. One type, associated with a positive ROA band at approximately 1300 cm(-1), dominates in hydrophobic environments and appears to be unhydrated; the other, associated with a positive ROA band at approximately 1340 cm(-1), dominates in hydrophilic environments and appears to be hydrated. Evidence is presented to support the hypothesis that unhydrated alpha-helix corresponds to the canonical conformation alpha(c) and hydrated alpha-helix to a more open conformation alpha(o) stabilized by hydrogen bonding of a water molecule or a hydrophilic side chain to the peptide carbonyl. Alpha-helical poly(L-lysine) and poly(L-ornithine) in aqueous solution and poly(L-alanine) in dichloracetic acid display both bands, but alpha-helical poly(l-glutamic acid) in aqueous solution and poly(gamma-benzyl L-glutamate) in CHCl(3) display only the approximately 1340 cm(-1) band and so may exist purely as alpha(o) due to enhanced stabilization of this conformation by particular side chain characteristics. The ROA spectrum of poly(beta-benzyl L-aspartate) in CHCl(3) reveals that it exists in a single left-handed alpha-helical state more analogous to alpha(o) than to alpha(c).  相似文献   

5.
Carbon-13 chemical shift anisotropy (CSA) tensors for various carbon sites of polypeptides, and for carbon sites in alpha-helical and beta-sheet conformations of poly-L-alanine, and polyglycine, are presented. The carbonyl (13)C CSA tensors were determined from one-dimensional CPMAS spectra obtained at a slow spinning speed, whereas the CSA tensors of C(alpha) and other carbons in side chains of peptides were determined using 2D PASS experiments on powder samples. The results suggest that the spans of (13)Carbonyl CSA tensors of alanine and glycine residues in various peptides are similar, even though the magnitude of individual components of the CSA tensor and the isotropic chemical shift are different. In addition, the delta(22) element is the only component of the (13)Carbonyl CSA tensor that significantly depends on the CO.HN hydrogen-bond length. Solid-state NMR experimental results also suggest that (13)Carbonyl and (13)C(alpha) CSA tensors are similar for alpha-helical and beta-sheet conformations of poly-L-alanine, which is in agreement with the reported quantum chemical calculation studies and previous solid-state NMR experimental studies on other systems. On the other hand, the (13)C(alpha) CSA tensor of the first alanine residue is entirely different from that of the second or later alanine residues of the peptide. While no clear trends in terms of the span and the anisotropic parameter were predicted for (13)C(beta) CSA tensors of alanine, they mainly depend on the conformation and dynamics of the side chain as well as on the packing interactions in the solid state of peptides.  相似文献   

6.
The structural characterization in crystals of three designed decapeptides containing a double d-segment at the C-terminus is described. The crystal structures of the peptides Boc-Leu-Aib-Val-Xxx-Leu-Aib-Val-(D)Ala-(D)Leu-Aib-OMe, (Xxx = Gly 2, (D)Ala 3, Aib 4) have been determined and compared with those reported earlier for peptide 1 (Xxx = Ala) and the all l analogue Boc-Leu-Aib-Val-Ala-Leu-Aib-Val-Ala-Leu-Aib-OMe, which yielded a perfect right-handed alpha-helical structure. Peptides 1 and 2 reveal a right-handed helical segment spanning residues 1 to 7, ending in a Schellman motif with (D)Ala(8) functioning as the terminating residue. Polypeptide chain reversal occurs at residue 9, a novel feature that appears to be the consequence of a C-H.O hydrogen bond between residue 4 C(alpha)H and residue 9 CO groups. The structures of peptides 3 and 4, which lack the pro R hydrogen at the C(alpha) atom of residue 4, are dramatically different. Peptide 3 adopts a right-handed helical conformation over the 1 to 7 segment. Residues 8 and 9 adopt alpha(L) conformations forming a C-terminus type I' beta-turn, corresponding to an incipient left-handed twist of the polypeptide chain. In peptide 4, helix termination occurs at Aib(6), with residues 6 to 9 forming a left-handed helix, resulting in a structure that accommodates direct fusion of two helical segments of opposite twist. Peptides 3 and 4 provide examples of chiral residues occurring in the less favored sense of helical twist; (D)Ala(4) in peptide 3 adopts an alpha(R) conformation, while (L)Val(7) in 4 adopts an alpha(L) conformation. The structural comparison of the decapeptides reported here provides evidence for the role of specific C-H.O hydrogen bonds in stabilizing chain reversals at helix termini, which may be relevant in aligning contiguous helical and strand segments in polypeptide structures.  相似文献   

7.
The conformational transition of hydrophobically end-capped poly(ethylene oxide), HP-PEO-HP [hydrophobic-poly(ethylene oxide)-hydrophobic], was studied using X-ray diffraction (XRD), differential scanning calorimetry (DSC), and Fourier transform infrared spectroscopy (FTIR) methods. Conformational transitions of HP-PEO-HP from a planar zigzag to a 7/2 helical conformation were observed as the molecular weight of the PEO main chain increased. HP-PEO-HP 1(18), with a PEO molecular weight of 1000 and 18 hydrocarbons on each end, has mainly an alpha-helical structure in poor solvents, whereas alpha and beta conformations coexist in good solvents. This means that the alpha-helical structure caused by the hydrogen bonds between the urethane linkages was broken by the high chain mobility caused by the melted adjacent chains of PEO, and instead, the beta-sheet was formed by the interaction of multiple hydrogen bonds. Another indication of hydrogen bonds breaking at increasing temperature is the transition of the N-H stretching peak in the FTIR data. HP-PEO-HP 2(18) and 4(18), which have 18 hydrocarbons on each end and PEO molecular weights of 2000 and 4000, respectively, and consist mostly of PEO, showed spherulites. This result also suggests that the PEO molecule has a 7/2 zigzag helical conformation. In contrast, HP-PEO-HP 1(18), which is composed of less PEO than HP-PEO-HP 2(18) and 4(18), did not show a spherulite structure.  相似文献   

8.
The solution conformations of several D,L copoly benzyl glutamates both random and alternating are studied by comparing their NMR spectra in chloroform and also in dioxane and dimethylformamide. The alpha CH chemical shifts characteristic of the alpha helix of strictly alternating DL copolymers in chloroform/0.5% TFA are established (3.65 and 3.82 ppm) and differ from that of the regular alpha helix (3.92 ppm). It is concluded that alternating copolymers prepared by an essentially racemization-free method are completely in the alpha-helical conformation which is characteristic of strictly alternating D-L copolymers, whereas random copolymers are largely regular alpha. The alpha leads to piDL helix/helix transition of an alternating copolymer in dioxane has been monitored and the alpha CH resonance characteristic of the piDL helix is found to be at the unusually low chemical shift of 4.45 ppm.  相似文献   

9.
The HIV-1 integrase (IN) catalyzes the integration of viral DNA in the human genome. In vitro the enzyme displays an equilibrium of monomers, dimers, tetramers and larger oligomers. However, its functional oligomeric form in vivo is not known. We report a study of the auto-associative properties of three peptides denoted K156, E156 and E159. These derive from the alpha4 helix of the IN catalytic core. The alpha4 helix is an amphipatic helix exposed at the surface of the protein and could be involved in the oligomerization process through its hydrophobic face. The peptides were obtained from the replacement of several amino acid residues by more helicogenic ones in the alpha4 helix peptide. K156 carries the basic residues Lys156 and Lys159, which have been shown important for the binding of IN to viral DNA. In E156 and E159 they are replaced with the acidic residue Glu. A fourth peptide K(E)156 obtained from the replacement of hydrophobic residues with Glu in K156 in order to abolish the auto-associative properties is used as a negative control. The capacity shown by peptides for alpha-helical formation is demonstrated by circular dichroism (CD) analysis performed in aqueous solution and in aqueous trifluoroethanol (TFE) mixtures. Both electrospray ionization mass spectrometry (ESI-MS) and glutaraldehyde chemical cross-linking show that peptides adopt different solvent-dependent equilibriums of monomers, dimers, trimers and tetramers. Oligomerization of peptides in aqueous solution is related to their ability to form helical structures. Addition of a small amount of TFE (<10%) stimulates helix stabilization and the interhelical hydrophobic contacts. Higher amounts of TFE alter the hydrophobic contacts and disrupt the oligomeric species. In addition to hydrophobic interactions, the patterns indicate that the biologically important Lys156 and Lys159 residues also participate in helix association. K(E)156 despite its ability to adopt a helical structure is unable to associate into oligomers, demonstrating the importance of hydrophobic contacts for oligomerization. Thus, the designed peptides provide us information on the functional properties of the alpha4 IN that seems to hold a dual role in DNA recognition and protein oligomerization.  相似文献   

10.
Tropomyosins from rabbit skeletal, chicken gizzard and earthworm muscle all exist as dimeric, ca. 100% alpha-helical coiled-coil species in benign media. Two major tropomyosin isoforms from each muscle source have been identified and can be conveniently designated alpha (fast) and beta (slow) based on electrophoretic mobility under denaturing conditions. The ratio of alpha to beta chains is ca. 3-4:1 for rabbit skeletal and ca. 1:1 for chicken gizzard and earthworm tropomyosins. Each chain from the former two muscle sources has been sequenced, thus providing a molecular basis for interpreting the in vivo population of homo- and hetero-dimers. The characteristics of each purified tropomyosin in weak-anion exchange, strong-cation exchange and reversed-phase high-performance liquid chromatography are described. Binding to and/or elution from the reversed-phase matrix results in dissociation into highly helical monomeric chains. This mode of chromatography separates the alpha and beta chains of earthworm and chicken gizzard tropomyosins, but not those of the rabbit protein. Both anion- and cation-exchange chromatography use mild (benign) elution conditions under which the native, in vivo dimer population should be preserved. Only the rabbit protein exhibited peak separation on the anion-exchange resin, with peak assignment corresponding to the known molecular organization of homo- and hetero-dimers. In strong cation-exchange analysis, all three tropomyosins exhibit a chromatographic transition near pH 6.5, possibly the result of histidine(s) titration. Collectively, the chromatographic data confirm the present understanding of the in vivo mixture of dimers for tropomyosin from rabbit skeletal and chicken gizzard. It is concluded that native earthworm tropomyosin exists predominantly as an alpha beta hetero-dimer.  相似文献   

11.
New molecules incorporating a uracil nucleic base and a dihydroindolizine (DHI) unit linked via spacer arms, i.e., uracil-spyrodihydroindolizine (4), were synthesised as models for light sensitive systems for nucleic acid labelling. The uracil-DHI (4) undergoes easy photocoloration to the uracil-betaine (5) generating the UV-detectable species. Preliminary results show rather weak binding of uracil-DHI (4) to calf thymus-DNA.  相似文献   

12.
Water-soluble porphyrins containing four platinum(II) complexes per molecule, [5alpha,10beta,15alpha,20beta-tetrakis(2-trans-(alpha,beta,alpha,beta-trans-Pt) and cis-(alpha,beta,alpha,beta-cis-Pt) [PtCl(NH(3))(2)]N-2-aminophenylporphyrin)], were synthesized and characterized. The binding of synthetic nucleotide polymers (poly(dG)-poly(dC), poly(dA)-poly(dT)) to the porphyrins was examined spectrophotometrically in aqueous solution. UV-vis spectral data suggested that these porphyrins bind to the nucleic acids by coordinative and Coulomb interactions.  相似文献   

13.
14.
A novel strategy for the synthesis of (R)- and (S)-alpha-methyl(alkyl)serine-containing peptides is presented. Using (S)-phenylalanine cyclohexylamide 6 as chiral auxiliary, the optically pure azlactones (R)- and (S)-2 were synthesized via a novel azlactone/oxazoline interconversion reaction (Figures 3 and 6). These azlactones constitute fully protected and activated synthetic equivalents of (R)- and (S)-alpha-methylserine and can be directly incorporated into peptides without further protective group manipulations. Like other alpha,alpha-dialkylated glycines, optically pure alpha-alkylserines can be used to stabilize beta-turn and alpha-helical conformations in short peptides.  相似文献   

15.
Four types of polypeptide (1)J(C alpha X) couplings are examined, involving the main-chain carbon C(alpha) and either of four possible substituents. A total 3105 values of (1)J(C alpha H alpha), (1)J(C alpha C beta), (1)J(C alpha C'), and (1)J(C alpha N') were collected from six proteins, averaging 143.4 +/- 3.3, 34.9 +/- 2.5, 52.6 +/- 0.9, and 10.7 +/- 1.2 Hz, respectively. Analysis of variances (ANOVA) reveals a variety of factors impacting on (1)J and ranks their relative statistical significance and importance to biomolecular NMR structure refinement. Accordingly, the spread in the (1)J values is attributed, in equal proportions, to amino-acid specific substituent patterns and to polypeptide-chain geometry, specifically torsions phi, psi, and chi(1) circumjacent to C(alpha). The (1)J coupling constants correlate with protein secondary structure. For alpha-helical phi, psi combinations, (1)J(C alpha H alpha) is elevated by more than one standard deviation (147.8 Hz), while both (1)J(C alpha N') and (1)J(C alpha C beta) fall short of their grand means (9.5 and 33.7 Hz). Rare positive phi torsion angles in proteins exhibit concomitant small (1)J(C alpha H alpha) and (1)J(C alpha N') (138.4 and 9.6 Hz) and large (1)J(C alpha C beta) (39.9 Hz) values. The (1)J(C alpha N') coupling varies monotonously over the phi torsion range typical of beta-sheet secondary structure and is largest (13.3 Hz) for phi around -160 degrees. All four coupling types depend on psi and thus help determine a torsion that is notoriously difficult to assess by traditional approaches using (3)J. Influences on (1)J stemming from protein secondary structure and other factors, such as amino-acid composition, are largely independent.  相似文献   

16.
The development of molecules that bind to specific protein surface sites and inhibit protein-protein interactions is a fundamental challenge in molecular recognition. New strategies for approaching this challenge could have important long-term ramifications in biology and medicine. We are exploring the concept that unnatural oligomers with well-defined conformations ("foldamers") can mimic protein secondary structural elements and thereby block specific protein-protein interactions. Here, we describe the identification and analysis of helical peptide-based foldamers that bind to a specific cleft on the anti-apoptotic protein Bcl-xL by mimicking an alpha-helical BH3 domain. Initial studies, employing a fluorescence polarization (FP) competition assay, revealed that among several alpha/beta- and beta-peptide foldamer backbones only alpha/beta-peptides intended to adopt 14/15-helical secondary structure display significant binding to Bcl-xL. The most tightly binding Bcl-xL ligands are chimeric oligomers in which an N-terminal alpha/beta-peptide segment is fused to a C-terminal alpha-peptide segment ((alpha/beta + alpha)-peptides)). Sequence-affinity relationships were probed via standard and nonstandard techniques (alanine scanning and hydrophile scanning, respectively), and the results allowed us to construct a computational model of the ligand/Bcl-xL complex. Analytical ultracentrifugation with a high-affinity (alpha/beta + alpha)-peptide established 1:1 ligand:Bcl-xL stoichiometry under FP assay conditions. Binding selectivity studies with the most potent (alpha/beta + alpha)-peptide, conducted via surface plasmon resonance measurements, revealed that this ligand binds tightly to Bcl-w as well as to Bcl-xL, while binding to Bcl-2 is somewhat weaker. No binding could be detected with Mcl-1. We show that our most potent (alpha/beta + alpha)-peptide can induce cytochrome C release from mitochondria, an early step in apoptosis, in cell lysates, and that this activity is dependent upon inhibition of protein-protein interactions involving Bcl-xL.  相似文献   

17.
Catalysis in the Julia-Colonna epoxidation of alpha,beta-unsaturated ketones is due to binding of the hydroperoxide enolate intermediate by the three N-terminal amidic N-H groups of alpha-helical poly-leucine; the N-terminal pair forms an oxy-anion hole, whilst the third aids displacement of hydroxide.  相似文献   

18.
We introduce a de novo designed peptide model system that enables the systematic study of 1) the role of a membrane environment in coiled-coil peptide folding, 2) the impact of different domains of an alpha-helical coiled-coil heptad repeat on the interaction with membranes, and 3) the dynamics of coiled-coil peptide-membrane interactions depending on environmental conditions. Starting from an ideal alpha-helical coiled-coil peptide sequence, several positively charged analogues were designed that exhibit a high propensity toward negatively charged lipid membranes. Furthermore, these peptides differ in their ability to form a stable alpha-helical coiled-coil structure. The influence of a membrane environment on peptide folding is studied. All positively charged peptides show strong interactions with negatively charged membranes. This interaction induces an alpha-helical structure of the former random-coil peptides, as revealed by circular dichroism measurements. Furthermore, vesicle aggregation is induced by a coiled-coil interaction of vesicle-bound peptides. Dynamic light scattering experiments show that the strength of vesicle aggregation increases with the peptide's intrinsic ability to form a stable alpha-helical coiled coil. Thus, the peptide variant equipped with the strongest inter- and intra-helical coiled-coil interactions shows the strongest effect on vesicle aggregation. The secondary structure of this peptide in the membrane-bound state was studied as well as its effect on the phospholipids. Peptide conformation within the peptide-lipid aggregates was analyzed by (13)C cross-polarization magic-angle spinning NMR experiments. A uniformly (13)C- and (15)N-labeled Leu residue was introduced at position 12 of the peptide chain. The (13)C chemical shift and torsion angle measurements support the finding of an alpha-helical structure of the peptide in its membrane-bound state. Neither membrane leakage nor fusion was observed upon peptide binding, which is unusual for amphiphatic peptide structures. Our results lay the foundation for a systematic study of the influence of the alpha-helical coiled-coil folding motif in membrane-active events on a molecular level.  相似文献   

19.
A method was developed for synthesizing alpha,alpha-disubstituted glycine residues bearing a large (more than 15-membered) hydrophobic ring. The ring-closing metathesis reactions of the dialkenylated malonate precursors proceed efficiently, particularly when long methylene chains tether both terminal olefin groups. Surprisingly, the amino groups of these alpha,alpha-disubstituted glycines are inert to conventional protective reactions (e.g., N-tert-butoxycarbonyl (Boc) protection: Boc(2)O/4-dimethylaminopyridine (DMAP)/CH(2)Cl(2); N-benzyloxycarbonyl (Z) protection: Z-Cl/DMAP/CH(2)Cl(2)). Curtius rearrangement of the carboxylic acid functionality of the malonate derivative after ring-closing metathesis leads to formation of an amine functionality and can be catalyzed by diphenylphosphoryl azide. However, only the intermediate isocyanates can be isolated, even in the presence of alcohols such as benzyl alcohol. The isocyanates obtained by Curtius rearrangement in an aprotic solvent (benzene) were isolated in high yields and treated with 9-fluorenylmethanol in a high-boiling-point solvent (toluene) under reflux to give the N-9-fluorenylmethoxycarbonyl (Fmoc)-protected aminomalonate derivatives in high yield. These hydrophobic amino acids can be incorporated into a peptide by Fmoc solid-phase peptide synthesis and the acid fluoride activation method. The stability of the monomeric alpha-helical structure of a 17-amino-acid peptide was enhanced by replacement of two alanine residues with two hydrophobic amino acid residues bearing a cyclic 18-membered ring. The results of sedimentation equilibrium studies suggested that the peptide assembles into hexamers in the presence of 100 mM NaCl.  相似文献   

20.
We synthesized two water-soluble porphyrins appending platinum(II) complexes [alpha,beta-(4a) and alpha,alpha-(4b) 5,15-bis(2-trans-[PtCl(NH3)2]N-2-aminoethylaminocarbonylphenyl) 2,3,7,8,12,13,17,18-octamethylporphyrin] and studied their reactions with a variety of nucleic acids [disodium adenosine-5'-monophosphate (AMP), disodium guanosine-5'-monophosphate (GMP), disodium thymidine-5'-monophosphate (TMP), disodium cytidine-5'-monophosphate (CMP), synthetic polymer poly(dG)-poly(dC), poly(dA)-poly(dT)] by 1H-NMR, UV-vis and FAB-MS spectroscopies. Based on the denaturation experiments of synthetic nucleic acid polymers, we conclude that the presence of the porphyrins (5.6 microM) does not cause significant changes in the melting temperature of poly(dA)-poly(dT) (28 microM) (deltaT=1 degrees C) and shows reannealing. On the other hand, gradual melting of poly(dG)-poly(dC) (28 microM) occurs at a low temperature (deltaT= -27 degrees C) in the presence of the porphyrins (5.6 microM), and the solutions do not show reannealing phenomena. The results of UV-vis and 1H-NMR experiments revealed that the porphyrins bind to guanine bases and that the porphyrins bind to GMP more strongly than to the other nucleotides. The binding modes between the porphyrins and synthetic nucleic acids are affected more by the coordination of the nucleobase [poly(dG)-poly(dC)] to the Pt(II) in the porphyrins than by Coulomb and hydrophobic interactions.  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