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 15 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.
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   

4.
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   

5.
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   

6.
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   

7.
Acrolein is an α,β‐unsaturated aldehyde that is a major environmental pollutant, as well as a product of cellular metabolism. DNA bases react with acrolein to form two regioisomeric exocyclic guanine adducts, namely γ‐hydroxy‐propanodeoxyguanosine (γ‐OH‐PdG) and its positional isomer α‐hydroxy‐propanodeoxyguanosine (α‐OH‐PdG). The γ‐OH‐PdG isomer adopts a ring‐opened conformation with minimal structural perturbation of the DNA host duplex. Conversely, the α‐OH‐PdG isomer assumes a ring‐closed conformation that significantly disrupts Watson‐Crick base‐pair alignments within the immediate vicinity of the damaged site. We have employed a combination of calorimetric and spectroscopic techniques to characterize the thermodynamic origins of these lesion‐induced structural alterations. Specifically, we have assessed the energetic impact of α‐OH‐PdG centered within an 11‐mer duplex by hybridizing the adduct‐containing oligonucleotide with its complementary strand harboring a central base N [where N = C or A], yielding a pair of duplexes containing the nascent lesion (α‐OH‐PdG·C) or mismatched adduct (α‐OH‐PdG·A), respectively. Our data reveal that the nascent lesion is highly destabilizing, whereas its mismatched counterpart partially ameliorates α‐OH‐PdG‐induced destabilization. Collectively, our data provide energetic characterizations of the driving forces that modulate error‐free versus error‐prone DNA translesion synthesis. The biological implications of our findings are discussed in terms of energetically probing acrolein‐mediated mutagenicity versus adduct‐induced genotoxicity. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 370–382, 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   

8.
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   

9.
Recent research has implicated the C‐terminus of G‐protein coupled receptors in key events such as receptor activation and subsequent intracellular sorting, yet obtaining structural information of the entire C‐tail has proven a formidable task. Here, a peptide corresponding to the full‐length C‐tail of the human CB1 receptor (residues 400–472) was expressed in E.coli and purified in a soluble form. Circular dichroism (CD) spectroscopy revealed that the peptide adopts an α‐helical conformation in negatively charged and zwitterionic detergents (48–51% and 36–38%, respectively), whereas it exhibited the CD signature of unordered structure at low concentration in aqueous solution. Interestingly, 27% helicity was displayed at high peptide concentration suggesting that self‐association induces helix formation in the absence of a membrane mimetic. NMR spectroscopy of the doubly labeled (15N‐ and 13C‐) C‐terminus in dodecylphosphocholine (DPC) identified two amphipathic α‐helical domains. The first domain, S401‐F412, corresponds to the helix 8 common to G protein‐coupled receptors while the second domain, A440‐M461, is a newly identified structural motif in the distal region of the carboxyl‐terminus of the receptor. Molecular modeling of the C‐tail in DPC indicates that both helices lie parallel to the plane of the membrane with their hydrophobic and hydrophilic faces poised for critical interactions. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 565–573, 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.
Atomic force microscopy has been used to follow in real time the adsorption from solution of two of the gliadin group of wheat seed storage proteins onto hydrophilic (mica) and hydrophobic (graphite) surfaces. The liquid cell of the microscope was used initially to acquire images of the substrate under a small quantity of pure solvent (1% acetic acid). Continuous imaging as an injection of gliadin solution entered the liquid cell enabled the adsorption process to be followed in situ from zero time. For ω‐gliadin, a monolayer was formed on the mica substrate during a period of ~2000 s, with the protein molecules oriented in parallel to the mica surface. In contrast, the ω‐gliadin had a relatively low affinity for the graphite substrate, as demonstrated by slow and weak adsorption to the surface. With γ‐gliadin, random deposition onto the mica surface was observed forming monodispersed structures, whereas on the graphite surface, monolayer islands of protein were formed with the protein molecules in a perpendicular orientation. Sequential adsorption experiments indicated strong interactions between the two proteins that, under certain circumstances, caused alterations to the surface morphologies of preadsorbed species. The results are relevant to our understanding of the interactions of proteins within the hydrated protein bodies of wheat grain and how these determine the processing properties of wheat gluten and dough. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 74–84, 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   

11.
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   

12.
Alkaline tropoelastin solutions (pH 11) were optically clear at low temperatures, but a firm gel formed when the temperature was raised to 37°C. Reversion to a clear solution took place if the temperature was lowered to below 20°C within less than 2 h, but not if 37°C was maintained for several hours. The precipitated elastin‐like hydrogel thus formed did not visually redissolve at low temperatures. Tropoelastin hydrogel was stable to subsequent washings with alkaline solution at 37°C, but at 4°C some hydrogel redissolved showing that association is at least partly reversible. Washing the hydrogel with neutral 8M urea solution at 4°C dissolved less than 10% of tropoelastin in 24 h. We characterized this phenomenon by combining temperature‐controlled light microscopy analysis, 1H NMR spectroscopy (temperature, diffusion, and relaxation time studies), and UV‐absorption‐based concentration measurements. The self‐association of tropoelastin at pH 11 is due to hydrophobic interactions in an emulsion‐like system in which the spherules coalesce in a manner like a water‐based latex paint that forms a durable hydrophobic sheet as water and the organic solvent evaporate. In the present case, the sedimentation and entanglement of the tropoelastin porous sheets means that reverse dissolution is a kinetically slow process. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 321–330, 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   

13.
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   

14.
The wheat gluten protein alpha-gliadin, a well known trigger of celiac disease, can be complexed by random copolymers of hydroxyethyl methacrylate (HEMA) and sodium 4-styrene sulfonate (SS). In this work, influence of alpha-gliadin and poly(HEMA-co-SS) concentrations on alpha-gliadin structure was studied using spectroscopic techniques and dynamic light scattering. In 70% ethanol or 0.06M HCl (pH 1.2), alpha-gliadin was found to self-associate upon increasing its concentrations and displayed decreased alpha-helical content and increased beta-turn and beta-sheet contents. At pH 1.2, alpha-gliadin interacted with poly(HEMA-co-SS) to form supra-molecular complex particles. Poly(HEMA-co-SS) induced alpha-gliadin structural changes that mimicked those obtained by varying the protein concentration in pure solution. At pH 6.8, alpha-gliadin was poorly soluble and formed large particles but alpha-helix is still main secondary structure. The influence of the polymer on protein structure was weaker at neutral than acidic pH. Interaction with poly(HEMA-co-SS) disrupted alpha-gliadin conformation and self-association to form new complex particles at neutral pH. This study provides insight into the mechanism of poly(HEMA-co-SS)/alpha-gliadin interaction and the polymer as alpha-gliadin sequestering agents in the supportive treatment of celiac disease.  相似文献   

15.
As a member of intrinsically unstructured protein family, β‐casein (β‐CN) contains relatively high amount of prolyl residues, adopts noncompact and flexible structure and exhibits chaperone‐like activity in vitro. Like many chaperones, native β‐CN does not contain cysteinyl residues and exhibits strong tendencies for self‐association. The chaperone‐like activities of three recombinant β‐CNs wild type (WT) β‐CN, C4 β‐CN (with cysteinyl residue in position 4) and C208 β‐CN (with cysteinyl residue in position 208), expressed and purified from E. coli, which, consequently, lack the phosphorylated residues, were examined and compared with that of native β‐CN using insulin and alcohol dehydrogenase as target/substrate proteins. The dimers (β‐CND) of C4‐β‐CN and C208 β‐CN were also studied and their chaperone‐like activities were compared with those of their monomeric forms. Lacking phosphorylation, WT β‐CN, C208 β‐CN, C4 β‐CN and C4 β‐CND exhibited significantly lower chaperone‐like activities than native β‐CN. Dimerization of C208 β‐CN with two distal hydrophilic domains considerably improved its chaperone‐like activity in comparison with its monomeric form. The obtained results demonstrate the significant role played by the polar contributions of phosphorylated residues and N‐terminal hydrophilic domain as important functional elements in enhancing the chaperone‐like activity of native β‐CN. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 623–632, 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   

16.
The N‐succinylamino acid racemases (NSAAR) belong to the enolase superfamily and they are large homooctameric/hexameric species that require a divalent metal ion for activity. We describe the structure and stability of NSAAR from Geobacillus kaustophilus (GkNSAAR) in the absence and in the presence of Co2+ by using hydrodynamic and spectroscopic techniques. The Co2+, among other assayed divalent ions, provides the maximal enzymatic activity at physiological pH. The protein seems to be a tetramer with a rather elongated shape, as shown by AU experiments; this is further supported by the modeled structure, which keeps intact the largest tetrameric oligomerization interfaces observed in other homooctameric members of the family, but it does not maintain the octameric oligomerization interfaces. The native functional structure is mainly formed by α‐helix, as suggested by FTIR and CD deconvoluted spectra, with similar percentages of structure to those observed in other protomers of the enolase superfamily. At low pH, the protein populates a molten‐globule‐like conformation. The GdmCl denaturation occurs through a monomeric intermediate, and thermal denaturation experiments indicate a high thermostability. The presence of the cofactor Co2+ did alter slightly the secondary structure, but it did not modify substantially the stability of the protein. Thus, GkNSAAR is one of the few members of the enolase family whose conformational propensities and stability have been extensively characterized. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 757–772, 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.
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   

18.
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   

19.
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   

20.
Among the pathological hallmarks of Alzheimer's disease (AD) is the deposition of amyloid‐β (Aβ) peptides, primarily Aβ (1–40) and Aβ (1–42), in the brain as senile plaques. A large body of evidence suggests that cognitive decline and dementia in AD patients arise from the formation of various aggregated forms of Aβ, including oligomers, protofibrils and fibrils. Hence, there is increasing interest in designing molecular agents that can impede the aggregation process and that can lead to the development of therapeutically viable compounds. Here, we demonstrate the ability of the specifically designed α,β‐dehydroalanine (ΔAla)‐containing peptides P1 (K‐L‐V‐F‐ΔA‐I‐ΔA) and P2 (K‐F‐ΔA‐ΔA‐ΔA‐F) to inhibit Aβ (1–42) aggregation. The mechanism of interaction of the two peptides with Aβ (1–42) seemed to be different and distinct. Overall, the data reveal a novel application of ΔAla‐containing peptides as tools to disrupt Aβ aggregation that may lead to the development of anti‐amyloid therapies not only for AD but also for many other protein misfolding diseases. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 456–465, 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   

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