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.
Acrolein, a cell metabolic product and main component of cigarette smoke, reacts with DNA generating α‐OH‐PdG lesions, which have the ability to pair with dATP during replication thereby causing G to T transversions. We describe the solution structure of an 11‐mer DNA duplex containing the mutagenic α‐OH‐PdG·dA base pair intermediate, as determined by solution nuclear magnetic resonance (NMR) spectroscopy and retrained molecular dynamics (MD) simulations. The NMR data support a mostly regular right‐handed helix that is only perturbed at its center by the presence of the lesion. Undamaged residues of the duplex are in anti orientation, forming standard Watson‐Crick base pairs alignments. Duplication of proton signals at and near the damaged base pair reveals the presence of two enantiomeric duplexes, thus establishing the exocyclic nature of the lesion. The α‐OH‐PdG adduct assumes a syn conformation pairing to its partner dA base that is protonated at pH 6.6. The three‐dimensional structure obtained by restrained molecular dynamics simulations show hydrogen bond interactions that stabilize α‐OH‐PdG in a syn conformation and across the lesion containing base pair. We discuss the implications of the structures for the mutagenic bypass of acrolein lesions. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 391–401, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

2.
A 34‐residue α/β peptide [IG(28–61)], derived from the C‐terminal part of the B3 domain of the immunoglobulin binding protein G from Streptoccocus, was studied using CD and NMR spectroscopy at various temperatures and by differential scanning calorimetry. It was found that the C‐terminal part (a 16‐residue‐long fragment) of this peptide, which corresponds to the sequence of the β‐hairpin in the native structure, forms structure similar to the β‐hairpin only at T = 313 K, and the structure is stabilized by non‐native long‐range hydrophobic interactions (Val47–Val59). On the other hand, the N‐terminal part of IG(28–61), which corresponds to the middle α‐helix in the native structure, is unstructured at low temperature (283 K) and forms an α‐helix‐like structure at 305 K, and only one helical turn is observed at 313 K. At all temperatures at which NMR experiments were performed (283, 305, and 313 K), we do not observe any long‐range connectivities which would have supported packing between the C‐terminal (β‐hairpin) and the N‐terminal (α‐helix) parts of the sequence. Such interactions are absent, in contrast to the folding pathway of the B domain of protein G, proposed recently by Kmiecik and Kolinski (Biophys J 2008, 94, 726–736), based on Monte‐Carlo dynamics studies. Alternative folding mechanisms are proposed and discussed. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 469–480, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

3.
Copolymers of sodium 4‐styrene sulfonate (SS) and hydroxyethyl methacrylate (HEMA) were investigated as sequestrants of α‐gliadin, a gluten protein, for the treatment of gluten intolerance. The interactions of α‐gliadin with poly(SS) and poly(HEMA‐co‐SS) with 9 and 26 mol% SS content were studied at gastric (1.2) and intestinal (6.8) pH using circular dichroism and measurements of turbidity, dynamic light scattering and zeta potential. The interactions and their influence on α‐gliadin secondary and aggregated structures depended mainly on the ratio of polymer negative and protein positive charges at pH 1.2, and on polymer SS content at polymer concentrations providing in excess of negative charges at either pH. Poly(SS) could not form complex particles with α‐gliadin in a sufficient excess of negative charges. Copolymerization with HEMA enhanced the formation of complex particles. Poly(HEMA‐co‐SS) with intermediate SS content was found to be the most effective sequestrant for α‐gliadin. This study provides insight into design considerations for polymer sequestrants used in the supportive treatment of celiac disease. © 2009 Wiley Periodicals, Inc. Biopolymers 93:418–428, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

4.
The modes of binding of 5′‐[4‐(aminoiminomethyl)phenyl]‐[2,2′‐Bifuran]‐5‐carboximidamide (DB832) to multi‐stranded DNAs: human telomere quadruplex, monomolecular R‐triplex, pyr/pur/pyr triplex consisting of 12 T*(T·A) triplets, and DNA double helical hairpin were studied. The optical adsorption of the ligand was used for monitoring the binding and for determination of the association constants and the numbers of binding sites. CD spectra of DB832 complexes with the oligonucleotides and the data on the energy transfer from DNA bases to the bound DB832 assisted in elucidating the binding modes. The affinity of DB832 to the studied multi‐stranded DNAs was found to be greater (Kass ≈ 107M?1) than to the duplex DNA (Kass ≈ 2 × 105M?1). A considerable stabilizing effect of DB832 binding on R‐triplex conformation was detected. The nature of the ligand tight binding differed for the studied multi‐stranded DNA depending on their specific conformational features: recombination‐type R‐triplex demonstrated the highest affinity for DB832 groove binding, while pyr/pur/pyr TTA triplex favored DB832 intercalation at the end stacking contacts and the human telomere quadruplex d[AG3(T2AG3)3] accommodated the ligand in a capping mode. Additionally, the pyr/pur/pyr TTA triplex and d[AG3(T2AG3)3] quadruplex bound DB832 into their grooves, though with a markedly lesser affinity. DB832 may be useful for discrimination of the multi‐sranded DNA conformations and for R‐triplex stabilization. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 8–20, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

5.
Modified internucleotide linkage featuring the C3′‐O‐P‐CH2‐O‐C4″ phosphonate grouping as an isosteric alternative to the phosphodiester C3′‐O‐P‐O‐CH2‐C4″ bond was studied in order to learn more on its stereochemical arrangement, which we showed earlier to be of prime importance for the properties of the respective oligonucleotide analogues. Two approaches were pursued: First, the attempt to prepare the model dinucleoside phosphonate with 13C‐labeled CH2 group present in the modified internucleotide linkage that would allow for a more detailed evaluation of the linkage conformation by NMR spectroscopy. Second, the use of ab initio calculations along with molecular dynamics (MD) simulations in order to observe the most populated conformations and specify main structural elements governing the conformational preferences. To deal with the former aim, a novel synthesis of key labeled reagent (CH3O)2P(O)13CH2OH for dimer preparation had to be elaborated using aqueous 13C‐formaldehyde. The results from both approaches were compared and found consistent. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 514–529, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

6.
Malondialdehyde induces frameshift mutations in Salmonella typhimurium strain hisD3052. The ability of propanodeoxyguanosine (PdG), a structural analog of the major malondialdehyde-deoxyguanosine adduct, to induce site-specific frameshift mutations was tested in the (CpG)4 hot-spot of hisD3052 carried on an M13 vector (M13MB102). PdG was introduced at position 6248 of duplex M13MB102 by ligation of the oligonucleotide 5'-CGC(PdG)CGGCATG-3' into a heteroduplex containing an 11-nucleotide gap in the (-)-strand between the SphI and BssHII restriction sites and deoxyuridine in place of thymidine in the (+)-strand. Ligation proceeded with 70% efficiency, and closed circular duplex DNA molecules were isolated in 40% yield. The adducted genome was sensitive to cleavage by SphI but resistant to cleavage by BssHII. Transformation of Escherichia coli strain JM105 with adducted M13MB102 led to 25% reduced survival relative to unadducted M13MB102 and produced frameshift mutations in 2.5% of the progeny phage. All of the mutations were deletions, and 70% occurred by deletion of CpG. Unadducted genomes exhibited a 40-fold lower mutation frequency, and all the mutations were single-base deletions at the sites of ligation of the 11-mer. These results illustrate that PdG, a structural analog of the major malondialdehyde-deoxyguanosine adduct, induces frameshift mutations in M13MB102 and that single-stranded nicks are efficient premutagenic lesions in this recombinant bacteriophage.  相似文献   

7.
Mucins form a group of heavily O‐glycosylated biologically important glycoproteins that are involved in a variety of biological functions, including modulating immune response, inflammation, and adhesion. Mucins are also involved in cancer and metastasis and often express diagnostic cancer antigens. Recently, a modified porcine submaxillary mucin (Tn‐PSM) containing GalNAcα1‐O‐Ser/Thr residues was shown to bind to soybean agglutinin (SBA) with ~106‐fold enhanced affinity relative to GalNAcα1‐O‐Ser, the pancarcinoma carbohydrate antigen. In this study, dynamic force spectroscopy is used to investigate molecular pairs of SBA and Tn‐PSM. A number of force jumps that demonstrate unbinding or rebinding events were observed up to a distance equal to 2.0 μm, consistent with the length of the mucin chain. The unbinding force increased from 103 to 402 pN with increasing force loading rate. The position of the activation barrier in the energy landscape of the interaction was 0.1 nm. The lifetime of the SBA–TnPSM complex in the absence of applied force was determined to be in the range 1.3–1.9 s. Kinetic parameters describing the rate of dissociation of other sugar lectin interactions are in the range 3.3 × 10?3–2.5 × 10?3 s. The long lifetime of the SBA‐TnPSM complex is compatible with a binding model in which lectin molecules “bind and jump” from α‐GalNAc residue to α‐GalNAc residue along the polypeptide chain of Tn‐PSM before dissociating. These findings have important implications for the molecular recognition properties of mucins. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 719–728, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

8.
Nucleic acid recognition is often mediated by α‐helices or disordered regions that fold into α‐helix on binding. A peptide bearing the DNA recognition helix of HPV16 E2 displays type II polyproline (PII) structure as judged by pH, temperature, and solvent effects on the CD spectra. NMR experiments indicate that the canonical α‐helix is stabilized at the N‐terminus, while the PII forms at the C‐terminus half of the peptide. Re‐examination of the dihedral angles of the DNA binding helix in the crystal structure and analysis of the NMR chemical shift indexes confirm that the N‐terminus half is a canonical α‐helix, while the C‐terminal half adopts a 310 helix structure. These regions precisely match two locally driven folding nucleii, which partake in the native hydrophobic core and modulate a conformational switch in the DNA binding helix. The peptide shows only weak and unspecific residual DNA binding, 104‐fold lower affinity, and 500‐fold lower discrimination capacity compared with the domain. Thus, the precise side chain conformation required for modulated and tight physiological binding by HPV E2 is largely determined by the noncanonical strained α‐helix conformation, “presented” by this unique architecture. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 432–443, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

9.
While end capping in α‐helices is well understood, the concept of capping a β‐hairpin is a relatively recent development; to date, favorable Coulombic interactions are the only example of sidechains at the termini influencing the overall stability of a β‐hairpin. While cross‐strand hydrophobic residues generally provide hairpin stabilization, particular when flanking the turn region, those remote from this location appear to provide little stabilization. While probing for an optimal residue at a hydrogen bond position near the terminus of a designed β‐hairpin a conservative, hydrophobic, V → I mutation was observed to not only result in a significant change in fold population but also effected major changes in the structuring shifts at numerous sites in the peptide. Mutational studies reveal that there is an interaction between the sidechain at this H‐bonded site and the sidechain at the C‐terminal non‐H‐bonded site of the hairpin. This interaction, which appears to be hydrophobic in character, requires a highly twisted hairpin structure. Modifications at the C‐terminal site, for example an E → A mutation (ΔΔGU = 6 kJ/mol), have profound affects on fold structure and stability. The data suggests that this may be a case of hairpin end capping by the formation of a hydrophobic cluster. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 557–564, 2009. This article was originally published online as an accepted preprint. The “Published Online”date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

10.
Exposure of cultured human melanocytes to ultraviolet radiation (UV) results in DNA damage. In melanoma, UV‐signature mutations resulting from unrepaired photoproducts are rare, suggesting the possible involvement of oxidative DNA damage in melanocyte malignant transformation. Here we present data demonstrating immediate dose‐dependent generation of hydrogen peroxide in UV‐irradiated melanocytes, which correlated directly with a decrease in catalase activity. Pretreatment of melanocytes with α‐melanocortin (α‐MSH) reduced the UV‐induced generation of 7,8‐dihydro‐8‐oxyguanine (8‐oxodG), a major form of oxidative DNA damage. Pretreatment with α‐MSH also increased the protein levels of catalase and ferritin. The effect of α‐MSH on 8‐oxodG induction was mediated by activation of the melanocortin 1 receptor (MC1R), as it was absent in melanocytes expressing loss‐of‐function MC1R, and blocked by concomitant treatment with an analog of agouti signaling protein (ASIP), ASIP‐YY. This study provides unequivocal evidence for induction of oxidative DNA damage by UV in human melanocytes and reduction of this damage by α‐MSH. Our data unravel some mechanisms by which α‐MSH protects melanocytes from oxidative DNA damage, which partially explain the strong association of loss‐of‐function MC1R with melanoma.  相似文献   

11.
Widely dispersed in genomic DNA, the tandem C‐rich repetitive stretches may fold below physiological pH, into i‐motif structures, stabilized by C·C+ pairing. Herein, structural status of a 9‐mer stretch d(CCCTAACCC), [the truncated double repeat of human telomeric sequence], and its extended version, comprising of additional ? TAA segment at the 3′‐end, representing the complete double repeat d(CCCTAACCCTAA), has been investigated. The pH dependent monophasic UV‐melting, Gel and CD data suggested that while the truncated version adopts a bimolecular i‐motif structure, its complete double repeat (12‐mer) sequence exists in two (bimolecular and tetramolecular) forms. A model is proposed for the tetramolecular i‐motif with conventional C · C+ base pairs, additionally stabilized by asymmetric A · A base pairs at the ?3′ TAA flanking ends and Watson–Crick A · T hydrogen bonding between intervening bases on antiparallel strands. Expanding the known topologies of DNA i‐motifs, such atypical geometries of i‐motifs may have implications in their recognition by proteins. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 150–160, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

12.
Zearalenone (ZEN) is a fusarotoxin converted predominantly into α‐zearalenol (α‐Zol) and β‐zearalenol (β‐Zol) by hepatic hydroxysteroid dehydrogenases. The feeding of naturally contaminated grains with ZEN was associated with hyperestrogenic and adverse effects on humans and animals. There is a lack of information on the attribution of the toxic effects of these toxins. One wonders if these effects are due to the parent molecule (ZEN) or to its major metabolites (α‐Zol and β‐Zol). Using human Caco‐2 cells, we looked for the molecular mechanisms of toxicity of ZEN, α‐Zol, and β‐Zol. Toxicity effects were studied by MTT viability assay and oxidative stress induction by measuring malondialdehyde (MDA) generation. To check whether the oxidative stress induction was associated to DNA lesions, we looked for DNA fragmentation by means of the Comet and the diphenylamine assays. To specify cell death pathway, we investigated caspase‐3 activation, confirmed by poly(ADP‐ribose) polymerase cleavage and by Bcl‐2 depletion. Our results clearly demonstrated that ZEN as well as its two metabolites presented variable toxic effects. They induced cell death and an increase in MDA generation. These effects were associated to DNA fragmentation as well as caspase‐3 activation. The observed toxic effects seem to be relieved by the metabolism of ZEN into α‐Zol and β‐Zol. © 2009 Wiley Periodicals, Inc. J Biochem Mol Toxicol 23:233–243, 2009; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/jbt.20284  相似文献   

13.
Young Kee Kang    Byung Jin Byun 《Biopolymers》2012,97(12):1018-1025
The γ‐peptide β‐turn structures have been designed computationally by the combination of chirospecific γ 2 , 3 ‐residues of 2‐(aminomethyl)cyclohexanecarboxylic acid (γAmc6) with a cyclohexyl constraint on the Cα?Cβ bond using density functional methods in water. The chirospecific γAmc6 dipeptide with the (2S,3S)‐(2R,3R) configurations forms a stable turn structure in water, resembling a type II′ turn of α‐peptides, which can be used as a β‐turn motif in β‐hairpins of Ala‐based α‐peptides. The γAmc6 dipeptide with homochiral (2S,3S)‐(2S,3S) configurations but different cyclohexyl puckerings shows the capability to be incorporated into one of two β‐turn motifs of gramicidin S. The overall structure of this gramicidin S analogue is quite similar to the native gramicidin S with the same patterns and geometries of hydrogen bonds. Our calculated results and the recently observed results may imply the wider applicability of chirospecific γ‐peptides with a cyclohexyl constraint on the backbone to form various peptide foldamers. © 2012 Wiley Periodicals, Inc. Biopolymers 97:1018–1025, 2012.  相似文献   

14.
The conformational preference and electronic properties of three L ‐tryptophyl‐containing dipeptides, i.e., glycyl‐L ‐tryptophane (H‐Gly‐Trp‐OH), L ‐alanyl‐L ‐tryptophane (H‐Ala‐Trp‐OH), and L ‐methionyl‐L ‐tryptophane (L ‐Met‐Trp‐OH) in solution depending on the pH of the media are studied both theoretically and experimentally. The effect of the protonation of the COO? and deprotonation of the NH as well as the alkaline hydrolysis of the amide fragment in a strong basic media on the electronic spectra are discussed. Ab initio and density functional theory (DFT) methods as well as the time‐dependent DFT (TD‐DFT) method as a function of the basis set are performed with a view to obtain the geometry and electronic properties of all of the species as well as the intermediate, obtained in the alkaline hydrolysis mechanism. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 727–734, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

15.
Melittin, the main hemolytic component of honeybee venom, is unfolded in an aqueous environment and folds into an α‐helical conformation in a lipid environment. Membrane fluidity is known to affect the activity and structure of melittin. By combining two structurally sensitive optical methods, circular dichroism (CD) and deep‐ultraviolet resonance Raman spectroscopy (dUVRR), we have identified distinct structural fluctuations in melittin correlated with increased and decreased 1,2‐dimyristoyl‐sn‐glycero‐3‐phosphocholine bilayer fluidities. CD spectra have reduced intensity at temperatures above 22°C and high concentrations of the cholesterol analog 5α‐cholestan‐3β‐ol indicating distortions in the α‐helical structure under these conditions. No increase in the amide S is observed in the temperature‐dependent dUVRR spectra, suggesting an increase in 310‐helical structure with increasing temperatures above 22°C. However, incorporation of 25 mol% 5α‐cholestan‐3β‐ol resulted in a small increase in the amide S intensity indicating partial unfolding of melittin. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 895–902, 2014.  相似文献   

16.
2SS[6‐127,64‐80] variant of lysozyme which has two disulfide bridges, Cys6‐Cys127 and Cys64‐Cys80, and lacks the other two disulfide bridges, Cys30‐Cys115 and Cys76‐Cys94, was quite unstructured in water, but a part of the polypeptide chain was gradually frozen into a native‐like conformation with increasing glycerol concentration. It was monitored from the protection factors of amide hydrogens against H/D exchange. In solution containing various concentrations of glycerol, H/D exchange reactions were carried out at pH* 3.0 and 4°C. Then, 1H‐15N‐HSQC spectra of partially deuterated protein were measured in a quenching buffer for H/D exchange (95% DMSO/5% D2O mixture at pH* 5.5 adjusted with dichloroacetate). In a solution of 10% glycerol, the protection factors were nearly equal to 10 at most of residues. With increasing glycerol concentration, some selected regions were further protected, and their protection factors reached about a 1000 in 30% glycerol solution. The highly protected residues were included in A‐, B‐, and C‐helices and β3‐strand, and especially centered on Ile 55 and Leu 56. In 2SS[6‐127,64‐80], long‐range interactions were recovered due to the preferential hydration by glycerol in the hydrophobic box of the α‐domain. Glycerol‐induced recovering of the native‐like structure is discussed from the viewpoint of molten globules growing with the protein folding. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 665–675, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

17.
Environmental factors, such as acidic pH, facilitate the assembly of α‐synuclein (α‐Syn) in aggregates, but the impact of pH on the very first step of α‐Syn aggregation remains elusive. Recently, we developed a single‐molecule approach that enabled us to measure directly the stability of α‐Syn dimers. Unlabeled α‐Syn monomers were immobilized on a substrate, and fluorophore‐labeled monomers were added to the solution to allow them to form dimers with immobilized α‐Syn monomers. The dimer lifetimes were measured directly from the fluorescence bursts on the time trajectories. Herein, we applied the single‐molecule tethered approach for probing of intermolecular interaction to characterize the effect of acidic pH on the lifetimes of α‐Syn dimers. The experiments were performed at pH 5 and 7 for wild‐type α?Syn and for two mutants containing familial type mutations E46K and A53T. We demonstrate that a decrease of pH resulted in more than threefold increase in the α‐Syn dimers lifetimes with some variability between the α‐Syn species. We hypothesize that the stabilization effect is explained by neutralization of residues 96–140 of α‐Syn and this electrostatic effect facilitates the association of the two monomers. Given that dimerization is the first step of α‐Syn aggregation, we posit that the electrostatic effect thereby contributes to accelerating α‐Syn aggregation at acidic pH. © 2016 Wiley Periodicals, Inc. Biopolymers 105: 715–724, 2016.  相似文献   

18.
Aminoglycoside phosphotransferase(3′)‐IIIa (APH) is the enzyme with broadest substrate range among the phosphotransferases that cause resistance to aminoglycoside antibiotics. In this study, the thermodynamic characterization of interactions of APH with its ligands are done by determining dissociation constants of enzyme–substrate complexes using electron paramagnetic resonance and fluorescence spectroscopy. Metal binding studies showed that three divalent cations bind to the apo‐enzyme with low affinity. In the presence of AMPPCP, binding of the divalent cations occurs with 7‐to‐37‐fold higher affinity to three additional sites dependent on the presence and absence of different aminoglycosides. Surprisingly, when both ligands, AMPPCP and aminoglycoside, are present, the number of high affinity metal binding sites is reduced to two with a 2‐fold increase in binding affinity. The presence of divalent cations, with or without aminoglycoside present, shows only a small effect (<3‐fold) on binding affinity of the nucleotide to the enzyme. The presence of metal–nucleotide, but not nucleotide alone, increases the binding affinity of aminoglycosides to APH. Replacement of magnesium (II) with manganese (II) lowered the catalytic rates significantly while affecting the substrate selectivity of the enzyme such that the aminoglycosides with 2′‐NH2 become better substrates (higher Vmax) than those with 2′‐OH. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 801–809, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

19.
Acrolein is a cell metabolic product and a main component of cigarette smoke. Its reaction with DNA produces two guanine lesions γ-OH-PdG, a major adduct that is nonmutagenic in mammalian cells, and the positional isomer α-OH-PdG. We describe here the solution structure of a short DNA duplex containing a single α-OH-PdG lesion, as determined by solution NMR spectroscopy and restrained molecular dynamics simulations. The spectroscopic data show a mostly regular right-handed helix, locally perturbed at its center by the presence of the lesion. All undamaged residues of the duplex are in anti orientation, forming standard Watson–Crick base-pair alignments. Duplication of proton signals near the damaged site differentiates two enantiomeric duplexes, thus establishing the exocyclic nature of the lesion. At the lesion site, α-OH-PdG rotates to a syn conformation, pairing to its counter cytosine residue that is protonated at pH 5.9. Three-dimensional models produced by restrained molecular dynamics simulations show different hydrogen-bonding patterns between the lesion and its cytosine partner and identify further stabilization of α-OH-PdG in a syn conformation by intra-residue hydrogen bonds. We compare the α-OH-PdG•dC duplex structure with that of duplexes containing the analogous lesion propano-dG and discuss the implications of our findings for the mutagenic bypass of acrolein lesions.  相似文献   

20.
Mei Zhang 《Biopolymers》2010,93(2):121-131
Recently, we isolated and purified a neutral polysaccharide (PGN) from edible fungus Pleurotus geestanus. Its structure was characterized by a range of physical–chemical methods, including high performance anion exchange chromatography, uronic acid, and protein analyses, size exclusion chromatography with ultraviolet, refractive index and light scattering detectors, and nuclear magnetic resonance. Our results revealed that PGN is a novel β‐(1→3)‐D ‐glucan with glucose attached to every other sugar residues at Position 6 in the backbone. It has a degree of branching of 1/2. Such structure is different from typical β‐(1→3)‐D ‐glucans schizophyllan and lentinan in which DB is 1/3 and 2/5, respectively. Rheological study showed a very interesting melting behavior of PGN in water solution: heating PGN in water leads to two transitions, in the range of 8–12.5°C and 25–60°C, respectively. The melting behavior and conformational changes were characterized by rheometry, micro‐differential scan calorimetry, atomic force microscopy, static and dynamic light scattering at different temperatures. The first heating‐induced transition corresponds to the disintegration of polymer bundles into small helical clusters, resembling the heating‐induced dissociation of SPG in water at 7°C; the second one might correspond to the dissociation of helical strands to individual chains. The ability of PGN to undergo a conformation/viscosity transition in water upon heating is very valuable to immobilize cells or enzymes or therapeutic DNA/RNA, which makes PGN a potentially useful biomaterial. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 121–131, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  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