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 989 毫秒
1.
2.
Small angle x-ray scattering was used to follow changes in the conformation and interactions of nucleosome core particles (NCP) as a function of the monovalent salt concentration Cs. The maximal extension (Dmax) of the NCP (145 ± 3-bp DNA) increases from 137 ± 5 Å to 165 ± 5 Å when Cs rises from 10 to 50 mM and remains constant with further increases of Cs up to 200 mM. In view of the very weak increase of the Rg value in the same Cs range, we attribute this Dmax variation to tail extension, a proposal confirmed by simulations of the entire I(q) curves, considering an ideal solution of particles with tails either condensed or extended. This tail extension is observed at higher salt values when particles contain longer DNA fragments (165 ± 10 bp). The maximal extension of the tails always coincides with the screening of repulsive interactions between particles. The second virial coefficient becomes smaller than the hard sphere virial coefficient and eventually becomes negative (net attractive interactions) for NCP145. Addition of salt simultaneously screens Coulombic repulsive interactions between NCP and Coulombic attractive interactions between tails and DNA inside the NCP. We discuss how the coupling of these two phenomena may be of biological relevance.  相似文献   

3.
We used small-angle X-ray solution scattering (SAXS) technique to investigate the nucleotide-mediated conformational changes of the head domains [subfragment 1 (S1)] of myosin V and VI processive motors that govern their directional preference for motility on actin. Recombinant myosin V-S1 with two IQ motifs (MV-S1IQ2) and myosin VI-S1 (MVI-S1) were engineered from Sf9 cells using a baculovirus expression system. The radii of gyration (Rg) of nucleotide-free MV-S1IQ2 and MVI-S1 were 48.6 and 48.8 Å, respectively. In the presence of ATP, the Rg value of MV-S1IQ2 decreased to 46.7 Å, while that of MVI-S1 increased to 51.7 Å, and the maximum chord length of the molecule decreased by ca 9% for MV-S1IQ2 and increased by ca 6% for MVI-S1. These opposite directional changes were consistent with those occurring in S1s with ADP and Vi or AlF4− 2 bound (i.e., in states mimicking the ADP/Pi-bound state of ATP hydrolysis). Binding of AMPPNP induced Rg changes of both constructs similar to those in the presence of ATP, suggesting that the timing of the structural changes for their motion on actin is upon binding of ATP (the pre-hydrolysis state) during the ATPase cycle. Binding of ADP to MV-S1IQ2 and MVI-S1 caused their Rg values to drop below those in the nucleotide-free state. Thus, upon the release of Pi, the reverse conformational change could occur, coupling to drive the directional motion on actin. The amount of Rg change upon the release of Pi was ca 6.4 times greater in MVI-S1 than in MV-S1IQ2, relating to the production of the large stroke of the MVI motor during its translocation on actin. Atomic structural models for these S1s based upon the ab initio shape reconstruction from X-ray scattering data were constructed, showing that MVI-S1 has the light-chain-binding domain positioned in the opposite direction to MV-S1IQ2 in both the pre- and post-powerstroke transition. The angular change between the light chain-binding domains of MV-S1IQ2 in the pre- to post-powerstroke transition was ∼ 50°, comparable to that of MII-S1. On the other hand, that of MVI-S1 was ∼ 100° or ∼ 130° much less than the currently postulated changes to allow the maximal stroke size of myosin VI-S1 but still significantly larger than those of other myosins reported so far. The results suggest that some additional alterations or elements are required for MVI-S1 to take maximal working strokes along the actin filament.  相似文献   

4.
We performed Raman and Brillouin scattering measurements to estimate glass transition temperature, Tg, of hydrated protein. The measurements reveal very broad glass transition in hydrated lysozyme with approximate Tg ∼ 180 ± 15 K. This result agrees with a broad range of Tg ∼ 160–200 K reported in literature for hydrated globular proteins and stresses the difference between behavior of hydrated biomolecules and simple glass-forming systems. Moreover, the main structural relaxation of the hydrated protein system that freezes at Tg ∼ 180 K remains unknown. We emphasize the difference between the “dynamic transition”, known as a sharp rise in mean-squared atomic displacement <r2> at temperatures around TD ∼ 200–230 K, and the glass transition. They have different physical origin and should not be confused.  相似文献   

5.
α-Crystallin, one of the main constituent proteins in the crystalline lens, is an important molecular chaperone both within and outside the lens. Presently, the structural relationship between α-crystallin and its target proteins during chaperone action is poorly understood. It has been hypothesised that target proteins bind within a central cavity. Small-angle neutron-scattering (SANS) experiments in conjunction with isotopic substitution were undertaken to investigate the interaction of a target lens protein (γE-crystallin) with α-crystallin (αH) and to measure the radius of gyration (Rg) of the proteins and their binary complexes in solution under thermal stress. The size of the αH in D2O incubated at 65 °C increased from 69 ± 3 to 81 ± 5 Å over 40 min, in good agreement with previously published small-angle X-ray scattering (SAXS) and SANS measurements. Deuterated γE-crystallin in H2O buffer (γED/H2O) and hydrogenous γE-crystallin in D2O buffer (γEH/D2O) free in solution were of insufficient size and/or too dilute to provide any measurable scattering over the angular range used, which was selected primarily to investigate γE:αH complexes. The evolution of the aggregation size/shape as an indicator of αH chaperone action was monitored by recording the neutron scattering in different H:D solvent contrasts under thermally stressed conditions (65 °C) for binary mixtures of αH, γEH, and γED. It was found that Rg(αH:γED/D2O) > Rg(αH:γEH/D2O) > Rg(αH/D2O) and that Rg(αH:γEH/D2O) ≈ Rg(αH/D2O). The relative sizes observed for the complexes weighted by the respective scattering powers of the various components imply that γE-crystallin binds in a central cavity of the α-crystallin oligomer, during chaperone action.  相似文献   

6.
A case study on Centaurea gymnocarpa Moris & De Not., a narrow endemic species, was carried out by analyzing its morphological, anatomical, and physiological traits in response to natural habitat stress factors under Mediterranean climate conditions. The results underline that the species is particularly adapted to the environment where it naturally grows. At the plant level, the above-ground/below-ground dry mass (1.73 ± 0.60) shows its investment predominately in the above-ground structure with a resulting total leaf area per plant of 1399 ± 94 cm2. The senescent attached leaves at the base of the plant contribute to limit leaf transpiration by shading soil around the plant. Moreover, the dense C. gymnocarpa leaf pubescence, leaf rolling, the relatively high leaf mass area (LMA = 12.3 ± 1.3 mg cm−2) and leaf tissue density (LTD = 427 ± 44 mg cm−3) contribute to limit leaf transpiration, also postponing leaf death under dry conditions. At the physiological level, a relatively low respiration/photosynthesis ratio (R/PN) in spring results from high R [2.26 ± 0.59 μmol (CO2) m−2 s−1] and PN [12.3 ± 1.5 μmol (CO2) m−2 s−1]. The high photosynthetic nitrogen use efficiency [PNUE = 15.5 ± 0.4 μmol (CO2) g−1 (N) s−1] shows the large amount of nitrogen (N) invested in the photosynthetic machinery of new leaves, associated to a high chlorophyll content (Chl = 35 ± 5 SPAD units). On the contrary, the highest R/PN ratio (1.75 ± 0.19) in summer is due to a significant PN decrease and increase of R in response to drought. The low PNUE [1.5 ± 0.2 μmol (CO2) g−1 (N) s−1] in this season is indicative of a greater N investment in leaf cell walls which may contribute to limit transpiration. On the contrary, the low R/PN ratio (0.05 ± 0.02) in winter is resulting from the limited enzyme activity of the respiratory apparatus [R = 0.23 ± 0.08 μmol (CO2) m−2 s−1] while the low PNUE [3.5 ± 0.2 μmol (CO2) g−1 (N) s−1] suggests that low temperatures additionally limit plant production. The experiment of the imposed water stress confirms that the C. gymnocarpa growth capability is in conformity with the severe conditions of its natural habitat, likewise as it may be the case with others narrow endemic species that have occupied niches with similar extreme conditions.  相似文献   

7.
Neuropathy target esterase (NTE) is an integral membrane protein localized in the endoplasmic reticulum in neurons. Irreversible inhibition of NTE by certain organophosphorus compounds produces a paralysis known as organophosphorus compound-induced delayed neuropathy. In vitro, NTE has phospholipase/lysophospholipase activity that hydrolyses exogenously added single-chain lysophospholipids in preference to dual-chain phospholipids, and NTE mutations have been associated with motor neuron disease. NTE's physiological role is not well understood, although recent studies suggest that it may control the cytotoxic accumulation of lysophospholipids in membranes. We used the NTE catalytic domain (NEST) to hydrolyze palmitoyl-2-hydroxy-sn-glycero-3-phosphocholine (p-lysoPC) to palmitic acid in bilayer membranes comprising 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) and the fluorophore 1-oleoyl-2-[12-[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl]-sn-glycero-3-phosphocholine (NBD-PC). Translational diffusion coefficients (DL) in supported bilayer membranes were measured by fluorescence recovery after pattern photobleaching (FRAPP). The average DL for DOPC/p-lysoPC membranes without NEST was 2.44 µm2s-1 ± 0.09; the DL for DOPC/p-lysoPC membranes containing NEST and diisopropylphosphorofluoridate, an inhibitor, was nearly identical at 2.45 ± 0.08. By contrast, the DL for membranes comprising NEST, DOPC, and p-lysoPC was 2.28 ± 0.07, significantly different from the system with inhibited NEST, due to NEST hydrolysis. Likewise, a system without NEST containing the amount of palmitic acid that would have been produced by NEST hydrolysis of p-lysoPC was identical at 2.26 ± 0.06. These results indicate that NTE's catalytic activity can alter membrane fluidity.  相似文献   

8.
As part of an effort to inhibit S100B, structures of pentamidine (Pnt) bound to Ca2+-loaded and Zn2+,Ca2+-loaded S100B were determined by X-ray crystallography at 2.15 Å (Rfree = 0.266) and 1.85 Å (Rfree = 0.243) resolution, respectively. These data were compared to X-ray structures solved in the absence of Pnt, including Ca2+-loaded S100B and Zn2+,Ca2+-loaded S100B determined here (1.88 Å; Rfree = 0.267). In the presence and absence of Zn2+, electron density corresponding to two Pnt molecules per S100B subunit was mapped for both drug-bound structures. One Pnt binding site (site 1) was adjacent to a p53 peptide binding site on S100B (± Zn2+), and the second Pnt molecule was mapped to the dimer interface (site 2; ± Zn2+) and in a pocket near residues that define the Zn2+ binding site on S100B. In addition, a conformational change in S100B was observed upon the addition of Zn2+ to Ca2+-S100B, which changed the conformation and orientation of Pnt bound to sites 1 and 2 of Pnt-Zn2+,Ca2+-S100B when compared to Pnt-Ca2+-S100B. That Pnt can adapt to this Zn2+-dependent conformational change was unexpected and provides a new mode for S100B inhibition by this drug. These data will be useful for developing novel inhibitors of both Ca2+- and Ca2+,Zn2+-bound S100B.  相似文献   

9.
In order to examine the effects of coordinated hydroxide ion and free hydroxide ion in configurational conversion of a tetraamine macrocyclic ligand complex, the kinetics of the cis-to-planar interconversion of cis-[Ni(isocyclam)(H2O)2]2+ (isocyclam, 1,4,7,11-tetraazacyclotetradecane) has been studied spectrophotometrically in basic aqueous solution. The interconversion requires the inversion of one sec-NH center of the folded cis-complex to have the planar species. Kinetic data are satisfactorily fitted by the rate law, R = kOH[OH][cis-[Ni(isocyclam)(H2O)2]2+], where kOH = 3.84 × 103 dm3 mol−1 s−1 at 25.0 ± 0.1 °C with I = 0.10 mol dm−3 (NaClO4). The large ΔH, 61.7 ± 3.2 kJ mol−1, and the large positive ΔS, 30.2 ± 10.8 J K−1 mol−1, strongly support a free-base-catalyzed mechanism for the reaction.  相似文献   

10.
Preparations, XPS and electronic spectroscopy, and magnetism of seven new one-dimensional cyano-bridged coordination polymers, chiral [Cu(RR-chxn)2][Pd(CN)4] · 2H2O (1), [Cu(trans-chxn)2][M(CN)4] · 2H2O (2, 4, and 6 for M = Pd, Ni, and Pt), and [Cu(cis-chxn)2][M(CN)4] · 2H2O (3, 5, and 7 for M = Pd, Ni, and Pt) (RR-chxn = cyclohexane-(1R,2R)-diamine, trans-chxn = racemic trans-cyclohexane-(1,2)-diamine, and cis-chxn = racemic cis-cyclohexane-(1,2)-diamine) have been reported in view of tuning of their electronic properties by stereochemistry of chxn ligands and metal-substitution. Comparison of Cu 2p1/2 and 2p3/2 peaks of XPS and broad d-d bands around 18 000 cm−1 of electronic spectra are described systematically for 1-7. Variable-temperature magnetic measurement shows that complexes 1-7 indicate weak antiferromagnetic interactions via cyano-bridges. Because of semi-coordination coupled with pseudo Jahn-Teller elongation and electrostatic interaction for 1, the axial Cu-N coordination bond distances of 2.330(7) and 3.092(8) Å are considerably longer than those of equatorial ones in the range from 2.016(6) to 2.030(6) Å. The former bond distances of 1 are intermediate values among the related Ni (2.324(6) and 3.120(8) Å) and Pt (2.34(1) and 3.09(1) Å) complexes.  相似文献   

11.
12.
This study uses low-angle (LAXS) and wide-angle (WAXS) X-ray synchrotron scattering, volume measurements and thin layer chromatography to determine the structure and interactions of SOPC, SOPC/cholesterol mixtures, SOPC/peptide and SOPC/cholesterol/peptide mixtures. N-acetyl-LWYIK-amide (LWYIK) represents the naturally-occurring CRAC motif segment in the pretransmembrane region of the gp41 protein of HIV-1, and N-acetyl-IWYIK-amide (IWYIK), an unnatural isomer, is used as a control. Both peptides thin the SOPC bilayer by ∼ 3 Å, and cause the area/unit cell (peptide + SOPC) to increase by ∼ 9 Å2 from the area/lipid of SOPC at 30 °C (67.0 ± 0.9 Å2). Model fitting suggests that LWYIK's average position is slightly closer to the bilayer center than IWYIK's, and both peptides are just inside of the phosphate headgroup. Both peptides increase the wide-angle spacing d of SOPC without cholesterol, whereas with 50% cholesterol LWYIK increases d but IWYIK decreases d. TLC shows that LWYIK is more hydrophobic than IWYIK; this difference persists in peptide/SOPC 1:9 mole ratio mixtures. Both peptides counteract the chain ordering effect of cholesterol to roughly the same degree, and both decrease KC, the bending modulus, thus increasing the SOPC membrane fluidity. Both peptides nucleate crystals of cholesterol, but the LWYIK-induced crystals are weaker and dissolve more easily.  相似文献   

13.
Complexes of the general formula cis-[MX2(PTA)2] (M = Pd, Pt; X = Cl, Br, I; PTA = 1,3,5-triaza-7-phosphaadamantane) were used to study the catalytic intramolecular hydroamination/cyclization of 4-pentyn-1-amine into 2-methyl-pyrroline in water, methanol, and dimethyl sulfoxide (DMSO). Kinetic data were measured via 1H NMR under homogeneous conditions at 50 °C and showed the following trends in rate: (i) Fastest rates were observed in D2O. (ii) The Pd complexes of this study produced faster rates than the Pt complexes. (iii) The identity of the halide had no effect on the catalytic rate. Cyclization by the catalytic precursor cis-[PdCl2(PTA)2] (4) in D2O was zero-order in substrate and first-order in metal complex with ΔH = 20.0 ± 2.1 kcal/mol, ΔS = −7.4 ± 6.3 cal/mol K, and Ea = 20.6 ± 2.1 kcal/mol. The acetylide complex, trans-[Pt(CC(CH2)3NH2)2(PTA)2] (6) precipitated from a catalytic mixture involving cis-[PtBr2(PTA)2] (2). Spectroscopic and kinetic studies indicated that 6 and its cis analog, 7, were the predominant species in solution and that they were both active catalysts for the cyclization reaction. These data, in conjunction with the rate trends, indicated that the mechanism of the Pd(II) and Pt(II) catalyzed hydroamination of terminal alkynylamines in aqueous solution followed a unique mechanism with cyclization of an acetylenic-amine ligand being rate determining.  相似文献   

14.
-crotamine is a small toxic protein (42 amino acid residues with three disulphide bridges) present in the venom of Crotallus durissus terrificus. Molecular parameters (R g=13.7 Å, S=3,000 Å2, V=9,200 Å3 and D max=40 Å) were derived from SAXS curves obtained from a solution of this protein at pH=4.5. An excellent agreement between the experimental distance distribution curve and that calculated from a model consisting of two lobes linked by the Cys(18)-Cys(30) disulphide bridge.  相似文献   

15.
Glycans comprise perhaps the largest biomass in nature, and more and more glycans are used in a number of applications, including those as pharmaceutical agents in the clinic. However, defining glycan molecular weight distributions during and after their preparation is not always straightforward. Here, we use pulse field gradient (PFG) 1H NMR self-diffusion measurements to assess molecular weight distributions in various glycan preparations. Initially, we derived diffusion coefficients, D, on a series of dextrans with reported weight-average molecular weights from about 5 kDa to 150 kDa. For each dextran sample, we analyzed 15 diffusion decay curves, one from each of the 15 major 1H resonance envelopes, to provide diffusion coefficients. By measuring D as a function of dextran concentration, we determined D at infinite dilution, Dinf, which allowed estimation of the hydrodynamic radius, Rh, using the Stokes-Einstein relationship. A plot of log Dinf versus log Rh was linear and provided a standard calibration curve from which Rh is estimated for other glycans. We then applied this methodology to investigate two other glycans, an α-(1→2)-l-rhamnosyl-α-(1→4)-d-galacturonosyl with quasi-randomly distributed, mostly terminal β(1→4)-linked galactose side-chains (GRG) and an α(1→6)-d-galacto-β(1→4)-d-mannan (Davanat), which is presently being tested against cancer in the clinic. Using the dextran-derived calibration curve, we find that average Rh values for GRG and Davanat are 76 ± 6 × 10−10 m and 56 ± 3 × 10−10 m, with GRG being more polydispersed than Davanat. Results from this study will be useful to investigators requiring knowledge of polysaccharide dispersity, needing to study polysaccharides under various solution conditions, or wanting to follow degradation of polysaccharides during production.  相似文献   

16.
In this study, we have structurally characterized the amidase of a nitrile-degrading bacterium, Rhodococcus sp. N-771 (RhAmidase). RhAmidase belongs to amidase signature (AS) family, a group of amidase families, and is responsible for the degradation of amides produced from nitriles by nitrile hydratase. Recombinant RhAmidase exists as a dimer of about 107 kDa. RhAmidase can hydrolyze acetamide, propionamide, acrylamide and benzamide with kcat/Km values of 1.14 ± 0.23 mM− 1s− 1, 4.54 ± 0.09 mM− 1s− 1, 0.087 ± 0.02 mM− 1s− 1 and 153.5 ± 7.1 mM− 1s− 1, respectively. The crystal structures of RhAmidase and its inactive mutant complex with benzamide (S195A/benzamide) were determined at resolutions of 2.17 Å and 2.32 Å, respectively. RhAmidase has three domains: an N-terminal α-helical domain, a small domain and a large domain. The N-terminal α-helical domain is not found in other AS family enzymes. This domain is involved in the formation of the dimer structure and, together with the small domain, forms a narrow substrate-binding tunnel. The large domain showed high structural similarities to those of other AS family enzymes. The Ser-cis Ser-Lys catalytic triad is located in the large domain. But the substrate-binding pocket of RhAmidase is relatively narrow, due to the presence of the helix α13 in the small domain. The hydrophobic residues from the small domain are involved in recognizing the substrate. The small domain likely participates in substrate recognition and is related to the difference of substrate specificities among the AS family amidases.  相似文献   

17.
New ruthenium(II) complexes with cyanamide ligands, cis-[Ru(bpy)2(Ipcyd)2] (1) and [Ru(bpy)2(OHpcyd)2] (2) (bpy = 2,2′-bipyridine, Ipcyd = 4-iodophenylcyanamide anion, OHpcyd = 4-(3-hydroxy-3-methylbut-1-ynil)phenylcyanamide), have been prepared and characterized by UV-Vis, IR and 1H NMR spectroscopies as well as electrochemical technique (CV). The complex cis-[Ru(bpy)2(Ipcyd)2] (1) crystallized with empirical formula of C34H24I2N8Ru in a monoclinic crystal system and space group of P21/c with a = 11.769(7) Å, b = 24.188(12) Å, c = 11.623(2) Å, β = 91.63(3)°, V = 3308(3) Å3 and Z = 4.  相似文献   

18.
A revision (C35r) to the CHARMM ether force field is shown to reproduce experimentally observed conformational populations of dimethoxyethane. Molecular dynamics simulations of 9, 18, 27, and 36-mers of polyethylene oxide (PEO) and 27-mers of polyethylene glycol (PEG) in water based on C35r yield a persistence length λ = 3.7 Å, in quantitative agreement with experimentally obtained values of 3.7 Å for PEO and 3.8 Å for PEG; agreement with experimental values for hydrodynamic radii of comparably sized PEG is also excellent. The exponent υ relating the radius of gyration and molecular weight of PEO from the simulations equals 0.515 ± 0.023, consistent with experimental observations that low molecular weight PEG behaves as an ideal chain. The shape anisotropy of hydrated PEO is 2.59:1.44:1.00. The dimension of the middle length for each of the polymers nearly equals the hydrodynamic radius Rh obtained from diffusion measurements in solution. This explains the correspondence of Rh and Rp, the pore radius of membrane channels: a polymer such as PEG diffuses with its long axis parallel to the membrane channel, and passes through the channel without substantial distortion.  相似文献   

19.
A Comparative Molecular Similarity Indices Analysis (CoMSIA) was performed for 2,6-substituted-4-monosubstituted aminopyrimidine antagonists of prostaglandin D2 receptor (DP). Both two-component (Q2 = 0.63, R2 = 0.82, SEE = 0.47 pIC50) and three-component (Q2 = 0.70, R2 = 0.91, SEE = 0.36 pIC50) CoMSIA models were established. Two hydrogen-bond acceptors with spatial separation of about 8 Å are shown as optimal for binding. A large hydrophobic center that separates the two acceptors confers to the potency of the 2,6-substituted-4-monosubstituted aminopyrimidine. The models were used to predict IC50 values for compounds which had functional groups different from those in the training set.  相似文献   

20.
Hideyuki Adachi  Isao Enami  Nobuo Kamiya 《BBA》2009,1787(2):121-128
Crystal structure of photosystem II (PSII) has been reported from prokaryotic cyanobacteria but not from any eukaryotes. In the present study, we improved the purification procedure of PSII dimers from an acidophilic, thermophilic red alga Cyanidium caldarium, and crystallized them in two forms under different crystallization conditions. One had a space group of P2221 with unit cell constants of a = 146.8 Å, b = 176.9 Å, and c = 353.7 Å, and the other one had a space group of P212121 with unit cell constants of a = 209.2 Å, b = 237.5 Å, and c = 299.8 Å. The unit cell constants of both crystals and the space group of the first-type crystals are different from those of cyanobacterial crystals, which may reflect the structural differences between the red algal and cyanobacterial PSII, as the former contains a fourth extrinsic protein of 20 kDa. X-ray diffraction data were collected and processed to a 3.8 Å resolution with the first type crystal. For the second type crystal, a post-crystallization treatment of dehydration was employed to improve the resolution, resulting in a diffraction data of 3.5 Å resolution. Analysis of this type of crystal revealed that there are 2 PSII dimers in each asymmetric unit, giving rise to 16 PSII monomers in each unit cell, which contrasts to 4 dimers per unit cell in cyanobacterial crystals. The molecular packing of PSII within the unit cell was constructed with the molecular replacement method and compared with that of the cyanobacterial crystals.  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