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.
2.
The human immunoglobulin G (IgG) class is the most prevalent antibody in serum, with the IgG1 subclass being the most abundant. IgG1 is composed of two Fab regions connected to a Fc region through a 15-residue hinge peptide. Two glycan chains are conserved in the Fc region in IgG; however, their importance for the structure of intact IgG1 has remained unclear. Here, we subjected glycosylated and deglycosylated monoclonal human IgG1 (designated as A33) to a comparative multidisciplinary structural study of both forms. After deglycosylation using peptide:N-glycosidase F, analytical ultracentrifugation showed that IgG1 remained monomeric and the sedimentation coefficients s020,w of IgG1 decreased from 6.45 S by 0.16–0.27 S. This change was attributed to the reduction in mass after glycan removal. X-ray and neutron scattering revealed changes in the Guinier structural parameters after deglycosylation. Although the radius of gyration (RG) was unchanged, the cross-sectional radius of gyration (RXS-1) increased by 0.1 nm, and the commonly occurring distance peak M2 of the distance distribution curve P(r) increased by 0.4 nm. These changes revealed that the Fab-Fc separation in IgG1 was perturbed after deglycosylation. To explain these changes, atomistic scattering modeling based on Monte Carlo simulations resulted in 123,284 and 119,191 trial structures for glycosylated and deglycosylated IgG1 respectively. From these, 100 x-ray and neutron best-fit models were determined. For these, principal component analyses identified five groups of structural conformations that were different for glycosylated and deglycosylated IgG1. The Fc region in glycosylated IgG1 showed a restricted range of conformations relative to the Fab regions, whereas the Fc region in deglycosylated IgG1 showed a broader conformational spectrum. These more variable Fc conformations account for the loss of binding to the Fcγ receptor in deglycosylated IgG1.  相似文献   

3.
Structural changes of barnase during folding were investigated using time-resolved small-angle X-ray scattering (SAXS). The folding of barnase involves a burst-phase intermediate, sometimes designated as the denatured state under physiological conditions, Dphys, and a second hidden intermediate. Equilibrium SAXS measurements showed that the radius of gyration (Rg) of the guanidine unfolded state (U) is 26.9 ± 0.7 Å, which remains largely constant over a wide denaturant concentration range. Time-resolved SAXS measurements showed that the Rg value extrapolated from kinetic Rg data to time zero, Rg,0, is 24.3 ± 0.1 Å, which is smaller than that of U but which is expanded from that of folding intermediates of other proteins with similar chain lengths (19 Å). After the burst-phase change, a single-exponential reduction in Rg2 was observed, which corresponds to the formation of the native state for the major component containing the native trans proline isomer. We estimated Rg of the minor component of Dphys containing the non-native cis proline isomer (Dphys,cis) to be 25.7 ± 0.6 Å. Moreover, Rg of the major component of Dphys containing the native proline isomer (Dphys,tra) was estimated as 23.9 ± 0.2 Å based on Rg,0. Consequently, both components of the burst-phase intermediate of barnase (Dphys,tra and Dphys,cis) are still largely expanded. It was inferred that Dphys possesses the N-terminal helix and the center of the β-sheet formed independently and that the formation of the remainder of the protein occurs in the slower phase.  相似文献   

4.
All-atom free-energy methods offer a promising alternative to kinetic molecular mechanics simulations of protein folding and association. Here we report an accurate, transferable all-atom biophysical force field (PFF02) that stabilizes the native conformation of a wide range of proteins as the global optimum of the free-energy landscape. For 32 proteins of the ROSETTA decoy set and six proteins that we have previously folded with PFF01, we find near-native conformations with an average backbone RMSD of 2.14 Å to the native conformation and an average Z-score of −3.46 to the corresponding decoy set. We used nonequilibrium sampling techniques starting from completely extended conformations to exhaustively sample the energy surface of three nonhomologous hairpin-peptides, a three-stranded β-sheet, the all-helical 40 amino-acid HIV accessory protein, and a zinc-finger ββα motif, and find near-native conformations for the minimal energy for each protein. Using a massively parallel evolutionary algorithm, we also obtain a near-native low-energy conformation for the 54 amino-acid engrailed homeodomain. Our force field thus stabilized near-native conformations for a total of 20 proteins of all structure classes with an average RMSD of only 3.06 Å to their respective experimental conformations.  相似文献   

5.
In all organisms adenylate kinases (Adks) play a vital role in cellular energy metabolism and nucleic acid synthesis. Due to differences in catalytic properties between the Adks found in prokaryotes and in the cytoplasm of eukaryotes, there is interest in targeting this enzyme for new drug therapies against infectious bacterial agents. Here we report the 2.1 Å resolution crystal structure for the 220-residue Adk from Burkholderia pseudomallei (BpAdk), the etiological agent responsible for the infectious disease melioidosis. The general structure of apo BpAdk is similar to other Adk structures, composed of a CORE subdomain with peripheral ATP-binding (ATPbd) and LID subdomains. The two molecules in the asymmetric unit have significantly different conformations, with a backbone RMSD of 1.46 Å. These two BpAdk conformations may represent ‘open’ Adk sub-states along the preferential pathway to the ‘closed’ substrate-bound state.  相似文献   

6.
Transmembrane proteins are embedded in cellular membranes of varied lipid composition and geometrical curvature. Here, we studied for the first time the allosteric effect of geometrical membrane curvature on transmembrane protein structure and function. We used single-channel optical analysis of the prototypic transmembrane β-barrel α-hemolysin (α-HL) reconstituted on immobilized single small unilamellar liposomes of different diameter and therefore curvature. Our data demonstrate that physiologically abundant geometrical membrane curvatures can enforce a dramatic allosteric regulation (1000-fold inhibition) of α-HL permeability. High membrane curvatures (1/diameter ∼1/40 nm−1) compressed the effective pore diameter of α-HL from 14.2 ± 0.8 Å to 11.4 ± 0.6 Å. This reduction in effective pore area (∼40%) when combined with the area compressibility of α-HL revealed an effective membrane tension of ∼50 mN/m and a curvature-imposed protein deformation energy of ∼7 kBT. Such substantial energies have been shown to conformationally activate, or unfold, β-barrel and α-helical transmembrane proteins, suggesting that membrane curvature could likely regulate allosterically the structure and function of transmembrane proteins in general.  相似文献   

7.
Glycosylation affects the physical properties of proteins in a number of ways including solubility and aggregation behavior. To elucidate the mechanism underlying these effects, we have measured second virial coefficients (A2) of the heavily glycosylated pheniophora lycii phytase (Phy) and its enzymatically deglycosylated counterpart (dgPhy) in native and in denatured form by means of small angle x-ray scattering. The measured A2-values show that the native forms of Phy and dgPhy are equally repulsive at the studied pH 8 where A2 equals 10.9 ± 0.1 × 104 mL mol g−2. However, when thermally denatured, the A2 of dgPhy decreases to 9.0 ± 0.2 × 104 mL mol g−2 whereas it remained unchanged for Phy. In accord with earlier investigations, the p(r)-function measured here suggested that the glycans did not affect the peptide structure of the native protein. Conversely, glycosylation markedly changed the structure of thermally denatured protein. This was evident from the radius of gyration, which increased by 32% for Phy and only 11% for dgPhy on denaturation. We suggest that this expanding effect of the glycans on the denatured protein conformation relies on steric hindrance that limits the range of torsion angles available to the polypeptide.  相似文献   

8.
9.
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   

10.
Bilayer liposomes from a mixture of dipalmitoylphosphatidylcholine (DPPC) and dipalmitoylphosphatidylethanolamine (DPPC:DPPE=8:2, molar ratio) or DPPC labeled with 14C-DPPC (DPPC:14C-DPPC) were bombarded with thermally activated tritium atoms. The tritiated liposomes were hydrolyzed by phospholipase C, and the tritium incorporation into different parts of the bilayer along its thickness was determined. The tritium flux attenuation coefficients were calculated for the headgroup (k1=0.176±0.032 Å–1) and acylglycerol residue (k2=0.046±0.004 Å–1) layers indicating a preferential attenuation of the tritium flux in the headgroup region and relative transparence of the membrane hydrophobic part. The finding is potentially important to apply tritium bombardment for investigation of spatial organization of transmembrane proteins in their native lipid environment.  相似文献   

11.
[Pt(Me2pipdt)2](BF4)2 salts [Me2pipdt = N,N-dimethyl-piperazine-2,3-dithione] bearing cationic dithiolene complexes react with (Bu4N)2[Pt(X)4] (X = SCN, CN) to form [Pt(Me2pipdt)2][Pt(SCN)4 ] (1) and [Pt(Me2pipdt)2][Pt(CN)4] (2) salts by metathesis. Black crystals of 1 have been structurally characterized showing that the two metals lie on inversion centers and exhibit a square planar coordination. The Pt-S bond distances in the anion complex (2.324(2) Å) are longer than in the cation complex (2.280(2) Å) whereas the C-S bond distances are shorter in SCN (average 1.669 Å) than in Me2pipdt (average 1.694 Å). The chelating Me2Pipdt ligand is found disordered in the λ/δ conformations with site occupancies of 50/50, respectively. The cation and anion complexes run parallel to a.  相似文献   

12.
DNA polymerase δ (Polδ) is a multisubunit polymerase that plays an indispensable role in replication from yeast to humans. Polδ from Saccharomyces cerevisiae is composed of three subunits: Pol3, Pol31, and Pol32. Despite the elucidation of the structures and models of the individual subunits (or portions, thereof), the nature of their assembly remains unclear. We present here a small-angle X-ray scattering analysis of a yeast Polδ complex (PolδT) composed of Pol3, Pol31, and Pol32N (amino acids 1-103 of Pol32). From the small angle X-ray scattering global parameters and reconstructed envelopes, we show that PolδT adopts an elongated conformation with a radius of gyration (Rg) of ∼ 52 Å and a maximal dimension of ∼ 190 Å. We also propose an orientation for the accessory Pol31-Pol32N subunits relative to the Pol3 catalytic core that best agrees with the experimental scattering profile. The analysis also points to significant conformational variability that may allow Polδ to better coordinate its action with other proteins at the replication fork.  相似文献   

13.
Synthesis and characterisation of the new macrocyclic ligand 1,7-dimethyl-4,10-di(methylcarbamoylmethy)-1,4,7,10-tetraazacyclododecane (L) are reported. The ligand, based on cyclen (1,4,7,10-tetraazacyclododecane), has been functionalised by the insertion of two methyl groups and two amidic pendant arms linked to the amine nitrogens. The interaction of L with H+, Na(I), Ca(II), Cu(II), Zn(II), Pb(II), and Gd(III) ions has been studied by potentiometric titrations, microcalorimetric and 1H NMR measurements in 0.1 mol dm−3 Me4NCl aqueous solution at 298.1±0.1 K. The thermodynamic data suggest that the N4 moiety is the binding site for Cu(II) and Zn(II), while in the case of Pb(II) also the pendant arms are coordinated to the metal ion. The crystal structure of [PbL](ClO4)2 (space group P21/a, a=12.883(2) Å, b=12.259(3) Å, c=17.275(5) Å, β=108.65(2)°, V=2585.0(11) Å3, Z=4, R=0.0660, RW 2=0.1467) shows the metal ion hexa-coordinated by the four nitrogen atoms of the cyclic tetra-amine and by the two amidic oxygens of the pendant arms.  相似文献   

14.
Previously reported crystal structures of free and DNA-bound dimers of λ Cro differ strongly (about 4 Å backbone rmsd), suggesting both flexibility of the dimer interface and induced-fit protein structure changes caused by sequence-specific DNA binding. Here, we present two crystal structures, in space groups P3221 and C2 at 1.35 and 1.40 Å resolution, respectively, of a variant of λ Cro with three mutations in its recognition helix (Q27P/A29S/K32Q, or PSQ for short). One dimer structure (P3221; PSQ form 1) resembles the DNA-bound wild-type Cro dimer (1.0 Å backbone rmsd), while the other (C2; PSQ form 2) resembles neither unbound (3.6 Å) nor bound (2.4 Å) wild-type Cro. Both PSQ form 2 and unbound wild-type dimer crystals have a similar interdimer β-sheet interaction between the β1 strands at the edges of the dimer. In the former, an infinite, open β-structure along one crystal axis results, while in the latter, a closed tetrameric barrel is formed. Neither the DNA-bound wild-type structure nor PSQ form 1 contains these interdimer interactions. We propose that β-sheet superstructures resulting from crystal contact interactions distort Cro dimers from their preferred solution conformation, which actually resembles the DNA-bound structure. These results highlight the remarkable flexibility of λ Cro but also suggest that sequence-specific DNA binding may not induce large changes in the protein structure.  相似文献   

15.
Small-angle neutron scattering (SANS) on the unilamellar vesicle (ULV) populations (diameter 500 and 1,000 Å) in D2O was used to characterize lipid vesicles from dimyristoylphosphatidylcholine (DMPC) at three phases: gel Lβ′, ripple Pβ′ and liquid Lα. Parameters of vesicle populations and internal structure of the DMPC bilayer were characterized on the basis of the separated form factor (SFF) model. Vesicle shape changes from nearly spherical in the Lα phase to elliptical in the Pβ′ and Lβ′ phases. This is true for vesicles prepared via extrusion through pores with the diameter 500 Å. Parameters of the internal bilayer structure (thickness of the membrane and the hydrophobic core, hydration and the surface area of the lipid molecule) were determined on the basis of the hydrophobic–hydrophilic (HH) approximation of neutron scattering length density across the bilayer ρ(x) and of the step function (SF) approximation of ρ(x). DMPC membrane thickness in the Lα phase (T=30°C) demonstrates a dependence on the membrane curvature for extruded vesicles. Prepared via extrusion through 500 Å diameter pores, vesicle population in the Lα phase has the following characteristics: average value of minor semi-axis 266±2 Å, ellipse eccentricity 1.11±0.02, polydispersity 26%, thickness of the membrane 48.9±0.2 Å and of the hydrophobic core 19.9±0.4 Å, surface area 60.7±0.5 Å2 and number of water molecules 12.8±0.3 per DMPC molecule. Vesicles prepared via extrusion through pores with the diameter 1,000 Å have polydispersity of 48% and membrane thickness of 45.5±0.6 Å in the Lα phase. SF approximation was used to describe the DMPC membrane structure in Lβ′ (T=10°C) and Pβ′ (T=20°C) phases. Extruded DMPC vesicles in D2O have membrane thickness of 49.6±0.5 Å in the Lβ′ phase and 48.3±0.6 Å in the Pβ′ phase. The dependence of the DMPC membrane thickness on temperature was restored from the SANS experiment.  相似文献   

16.
The thermostable direct hemolysin (TDH) is a major virulence factor of Vibrio parahaemolyticus. We have characterized the conformational properties of TDH by small-angle X-ray scattering (SAXS), ultracentrifugation and transmission electron microscopy. Sedimentation equilibrium and velocity studies revealed that the protein is tetrameric in aqueous solvents. The Guinier plot derived from SAXS data provided a radius of gyration of 29.0 Å. The elongated pattern with a shoulder of a pair distance distribution function derived from SAXS data suggested the presence of molecules with an anisotropic shape having a maximum diameter of 98 Å. Electron microscopic image analysis of the negatively stained TDH oligomer showed the presence of C4 symmetric particles with edge and diagonal lengths of 65 Å and 80 Å, respectively. Shape reconstruction was carried out by ab initio calculations using the SAXS data with a C4 symmetric approximation. These results suggested that the tetrameric TDH assumes an oblate structure. The hydrodynamic parameters predicted from the ab initio model differed slightly from the experimental values, suggesting the presence of flexible segments.  相似文献   

17.

Background

Serum albumin is a major pharmacokinetic effector of drugs. To gain further insight into albumin binding chemistry, the crystal structures of six oncology agents were determined in complex with human serum albumin at resolutions of 2.8 to 2.0 Å: camptothecin, 9-amino-camptothecin, etoposide, teniposide, bicalutamide and idarubicin.

Methods

Protein crystal growth and low temperature X-ray crystallography

Results

These large, complex drugs are all bound within the subdomain IB binding region which can be described as a hydrophobic groove formed by α-helices h7, h8 and h9 covered by the extended polypeptide L1. L1 creates a binding cavity with two access sites, one between loop L1 and α-helices h7 and h8 (distal site: IBd) and the other between L1 and α-helix h9 (proximal site: IBp). Camptothecin (2.4 Å) and 9 amino camptothecin (2.0 Å) are clearly bound as the open lactone form (IBp). Idarubicin (2.8 Å) binds in a DNA like dimer complex via an intermolecular π stacking arrangement in IBd. Bicalutamide (2.4 Å) is bound in a folded intramolecular π stacking arrangement between two aromatic rings in IBd similar to idarubicin. Teniposide (2.7 Å) and etoposide (2.7 Å), despite small chemical differences, are bound in two distinctly different sites at or near IB. Teniposide is internalized via primarily hydrophobic interactions and spans through both openings (IBp-d). Etoposide is bound between the exterior of IB and IIA and exhibits an extensive hydrogen bonding network.

Conclusions

Subdomain IB is a major binding site for complex heterocyclic molecules.

General significance

The structures have important implications for drug design and development. This article is part of a Special Issue entitled Serum Albumin.  相似文献   

18.
Dihydrodipicolinate synthase (DHDPS, E.C. 4.2.1.52), a validated antibiotic target, catalyses the first committed step in the lysine biosynthetic pathway: the condensation reaction between (S)-aspartate β-semialdehyde [(S)-ASA] and pyruvate via the formation of a Schiff base intermediate between pyruvate and the absolutely conserved active-site lysine. Escherichia coli DHDPS mutants K161A and K161R of the active-site lysine were characterised for the first time. Unexpectedly, the mutant enzymes were still catalytically active, albeit with a significant decrease in activity. The kcat values for DHDPS-K161A and DHDPS-K161R were 0.06 ± 0.02 s−1 and 0.16 ± 0.06 s−1 respectively, compared to 45 ± 3 s−1 for the wild-type enzyme. Remarkably, the KM values for pyruvate increased by only 3-fold for DHDPS-K161A and DHDPS-K161R (0.45 ± 0.04 mM and 0.57 ± 0.06 mM, compared to 0.15 ± 0.01 mM for the wild-type DHDPS), while the KM values for (S)-ASA remained the same for DHDPS-K161R (0.12 ± 0.01 mM) and increased by only 2-fold for DHDPS-K161A (0.23 ± 0.02 mM) and the Ki for lysine was unchanged. The X-ray crystal structures of DHDPS-K161A and DHDPS-K161R were solved at resolutions of 2.0 and 2.1 Å respectively and showed no changes in their secondary or tertiary structures when compared to the wild-type structure. The crystal structure of DHDPS-K161A with pyruvate bound at the active site was solved at a resolution of 2.3 Å and revealed a defined binding pocket for pyruvate that is thus not dependent upon lysine 161. Taken together with ITC and NMR data, it is concluded that although lysine 161 is important in the wild-type DHDPS-catalysed reaction, it is not absolutely essential for catalysis.  相似文献   

19.
Human complement receptor type 1 (CR1, CD35) is a type I membrane-bound glycoprotein that belongs to the regulators of complement activity (RCA) family. The extra-cellular component of CR1 is comprised of 30 short complement regulator (SCR) domains, whereas complement receptor type 2 (CR2) has 15 SCR domains and factor H (FH) has 20 SCR domains. The domain arrangement of a soluble form of CR1 (sCR1) was studied by X-ray scattering and analytical ultracentrifugation. The radius of gyration RG of sCR1 of 13.4(±1.1) nm is not much greater than those for CR2 and FH, and its RG/R0 anisotropy ratio is 3.76, compared to ratios of 3.67 for FH and 4.1 for CR2. Unlike CR2, but similar to FH, two cross-sectional RG ranges were identified that gave RXS values of 4.7(±0.2) nm and 1.2(±0.7) nm, respectively, showing that the SCR domains adopt a range of conformations including folded-back ones. The distance distribution function P(r) showed that the most commonly occurring distance in sCR1 is at 11.5 nm. Its maximum length of 55 nm is less than double those for CR2 or FH, even though sCR1 has twice the number of SCR domains compared to CR2 Sedimentation equilibrium experiments gave a mean molecular weight of 235 kDa for sCR1. This is consistent with the value of 245 kDa calculated from its composition including 14 N-linked oligosaccharide sites, and confirmed that sCR1 is a monomer in solution. Sedimentation velocity experiments gave a sedimentation coefficient of 5.8 S. From this, the frictional ratio (f/f0) of sCR1 was calculated to be 2.29, which is greater than those of 1.96 for CR2 and 1.77 for FH. The constrained scattering modelling of the sCR1 solution structure starting from homologous SCR domain structures generated 5000 trial conformationally randomised models, 43 of which gave good scattering fits to show that sCR1 has a partly folded-back structure. We conclude that the inter-SCR linkers show structural features in common with those in FH, but differ from those in CR2, and the SCR arrangement in CR1 will permit C3b or C4b to access all three ligand sites.  相似文献   

20.
The assignment of the 1H and 15Nnuclear magnetic resonance spectra of the Src-homology region 3 domain ofchicken brain -spectrin has been obtained. A set of solutionstructures has been determined from distance and dihedral angle restraints,which provide a reasonable representation of the protein structure insolution, as evaluated by a principal component analysis of the globalpairwise root-mean-square deviation (rmsd) in a large set of structuresconsisting of the refined and unrefined solution structures and the crystalstructure. The solution structure is well defined, with a lower degree ofconvergence between the structures in the loop regions than in the secondarystructure elements. The average pairwise rmsd between the 15 refinedsolution structures is 0.71 ± 0.13 Å for the backbone atoms and1.43 ± 0.14 Å for all heavy atoms. The solution structure isbasically the same as the crystal structure. The average rmsd between the 15refined solution structures and the crystal structure is 0.76 Å forthe backbone atoms and 1.45 ± 0.09 Å for all heavy atoms. Thereare, however, small differences probably caused by intermolecular contactsin the crystal structure.  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