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.
Recent studies have demonstrated the feasibility of using membrane ultrafiltration for the purification of pegylated proteins; however, the separations have all been performed at relatively low protein concentrations where intermolecular interactions are unimportant. The objective of this study was to examine the behavior at higher PEG concentrations and to develop an appropriate theoretical framework to describe the effects of intermolecular interactions. Ultrafiltration experiments were performed using pegylated α‐lactalbumin as a model protein with both neutral and charged composite regenerated cellulose membranes. The transmission of the pegylated α‐lactalbumin, PEG, and α‐lactalbumin all increase with increasing PEG concentration due to the increase in the solute partition coefficient arising from unfavorable intermolecular interactions in the bulk solution. The experimental results were in good agreement with a simple model that accounts for the change in Gibbs free energy associated with these intermolecular interactions, including the effects of concentration polarization on the local solute concentrations upstream of the membrane. These intermolecular interactions are shown to cause a greater than expected loss of pegylated product in a batch ultrafiltration system, and they alter the yield and purification factor that can be achieved during a diafiltration process to remove unreacted PEG. © 2013 American Institute of Chemical Engineers Biotechnol. Prog., 29:655–663, 2013  相似文献   

2.
The retroviral Gag polyprotein directs virus particle assembly, resulting in the release of virions from the plasma membranes of infected cells. The earliest steps in assembly, those immediately following Gag synthesis, are very poorly understood. For Rous sarcoma virus (RSV), Gag proteins are synthesized in the cytoplasm and then undergo transient nuclear trafficking before returning to the cytoplasm for transport to the plasma membrane. Thus, RSV provides a useful model to study the initial steps in assembly because the early and later stages are spatially separated by the nuclear envelope. We previously described mutants of RSV Gag that are defective in nuclear export, thereby isolating these “trapped” Gag proteins at an early assembly step. Using the nuclear export mutants, we asked whether Gag protein-protein interactions occur within the nucleus. Complementation experiments revealed that the wild-type Gag protein could partially rescue export-defective Gag mutants into virus-like particles (VLPs). Additionally, the export mutants had a trans-dominant negative effect on wild-type Gag, interfering with its release into VLPs. Confocal imaging of wild-type and mutant Gag proteins bearing different fluorescent tags suggested that complementation between Gag proteins occurred in the nucleus. Additional evidence for nuclear Gag-Gag interactions was obtained using fluorescence resonance energy transfer, and we found that the formation of intranuclear Gag complexes was dependent on the NC domain. Bimolecular fluorescence complementation allowed the direct visualization of intranuclear Gag-Gag dimers. Together, these experimental results strongly suggest that RSV Gag proteins are capable of interacting within the nucleus.  相似文献   

3.
By controlled dehydration, the unit cells of dodecamer DNA–drug crystals have been shrunk from 68 000 (normal state) to 60 000 (partially dehydrated intermediate state) to 51 000 Å3 (fully dehydrated state), beyond which no further solvent loss occurs. The total solvent content in the normal crystals is ~40% by volume, reducing to ~20% in the fully dehydrated phase. The 25% reduction in cell volume induced a dramatic enhancement in the resolution of the X-ray diffraction data (from 2.6 to beyond 1.5 Å). We have determined the structures of the normal, partially dehydrated and fully dehydrated crystals. Details of the ligand binding have been presented in the preceding article. The present paper describes the unique features of the structure of the fully dehydrated phase. This structure was refined with 9015 unique observed reflections to R = 14.9%, making it one of the most reliable models of B-form DNA available. The crystals exist as infinite polymeric networks, in which neighbouring dodecamer duplexes are crosslinked through phosphate oxygens via direct bonding to magnesium cations. The DNA is packed so tightly that there is essentially only a single layer of solvent between adjacent molecules. The details of the crystal packing, magnesium bridging, DNA hydration and DNA conformation are described and compared with other experimental evidence related to DNA condensation.  相似文献   

4.
2-[(2,5-dichloro-4-nitro-phenylamino)-methoxy-methyl]-8-hydroxy-quinoline 1 and 2-methyl-quinoline-5,8-dione-5-oxime 2 were obtained as potential HIV-1 integrase inhibitors and analyzed by X-ray crystallography. Semiempirical theoretical calculations of energy preferred conformations were also carried out. The crystal structures of both compounds are stabilized via hydrogen bonds and pi-pi stacking interactions. The planarity of compound 1 is caused by intramolecular hydrogen bonds.  相似文献   

5.
We have constructed three-dimensional models of four pathogenesis-related (PR) proteins from wheat (wheatwins) belonging to the PR-4 family. All the models were based on the knowledge of the tertiary structure of barwin, a highly homologous protein from barley. Wheatwin1 and wheatwin2 differ in two amino acid residues (positions 62 and 68) out of 125. Wheatwin4 differs from wheatwin2 in one residue at position 78, while wheatwin3 differs from wheatwin1 in one residue at position 88. The global folding and the secondary structures were very similar through all the sequences, including the regions of the amino acid substitutions. The main differences were found in the traits 15-21, 84-86 and 91-93. Trait 15-21 was predicted as ss-sheet in wheatwin4 and random-coil in the other proteins. Trait 84-86 was predicted as ss-sheet in wheatwin3 and random-coil in the other proteins. Trait 91-93 was predicted as random coil in wheatwin1 and wheatwin3 and ss-sheet in the other two proteins. Traits 15-21 and 84-86 were exposed, while trait 91-93 was quite hidden in all the proteins. The antifungal activities of the four proteins towards the specific pathogenic fungus Fusarium culmorum were distinct and well correlated to the structural differences. These results suggest that these regions may have a role in the action mechanism, which is still unknown.  相似文献   

6.
Most biopolymers exist in a plasticized state, whether it is naturally with water or unnaturally with glycerol or other suitable polyol, to make a flexible material. We have found that the extent to which a biopolymer can be plasticized is dependent on its molecular and higher order structures outside of simply molecular weight. Lactalbumin, ovalbumin, corn zein, wheat gluten, and feather keratin were plasticized with glycerol from very low to very high amounts. The conformation of the proteins was monitored with Fourier transform-infrared (FT-IR) spectroscopy and X-ray powder diffraction (XRD) and correlated with the tensile modulus. Protein conformational changes were pronounced for polar proteins with a low amount of cysteine. FT-IR showed that the conformational changes resulted in ordering of the protein at low to moderate plasticization levels. For proteins with little resistance to conformational changes, additional small-scale ordering occurred around the glass transition, as observed in XRD. Accurate comparison of plasticized proteins was dependent on knowing whether or not the protein was glassy or rubbery at room temperature as no differences arose in the glassy state. The transition from glassy to rubbery behavior with plasticization level can be found from modulus, FT-IR, and XRD data.  相似文献   

7.
8.
The mechanisms of intermolecular protein complex formation were studied by the example of monomers, oligomers and aggregates of bovine serum albumin (BSA) depending on the protein concentration, pH and urea concentration. Using dynamic light scattering (DLS), analytical ultracentrifugation (AUC) and PAG electrophoresis we have shown the existence of dynamic equilibrium between monomers and aggregates in BSA solution. Decreasing pH of the solution (4.0–1.0) resulted in increasing sizes of the aggregates. In the solutions with low urea concentrations (below 2 M) the sizes of aggregates decreased, while higher urea concentrations (2–8 M) induced formation of larger aggregates due to the unfolding of the protein.  相似文献   

9.
The temperature window for the extrusion of glycerol-plasticized wheat gluten was increased by the use of salicylic acid, a known scorch retarder and radical scavenger. It was possible to extrude 30 wt % glycerol-wheat gluten films with a die-head temperature as high as 135 degrees C, rather than 95 degrees C, by incorporating only 1 wt % salicylic acid. Small effects of shear-induced heating during extrusion at the higher temperatures suggested that the acid acted as a lubricant and viscosity reducer. The latter was suggested to originate primarily from the salicylic-acid-induced reduction in the degree of protein aggregation/cross-linking, as indicated by size-exclusion high-performance liquid chromatography and chemiluminescence. Electron paramagnetic resonance spectroscopy on extruded films indicated that the beneficial effect of salicylic acid was due to its radical scavenging effect. Tensile tests on extrudates revealed that the materials produced at the substantially higher processing temperature were still ductile. The complex shear modulus increased more slowly with increasing salicylic acid content above 110-120 degrees C, indicating that the aggregation/cross-linking rate was slower with salicylic acid, that is, that it did have a scorch-retarding effect, besides yielding a lower final degree/complexity of aggregation.  相似文献   

10.
《Biophysical journal》2022,121(15):2873-2881
Molecular interactions and reactions in living cells occur with high background concentrations of organic compounds including proteins. Uncharged water-soluble polymers are commonly used cosolutes in studies on molecular crowding, and most studies argue about the effects of intracellular crowding based on results obtained using polymer cosolutes. Further investigations using protein crowders and organic cations are important in understanding the effects of cellular environments on nucleic acids with negatively charged surfaces. We assessed the effects of using model globular proteins, serum proteins, histone proteins, structurally flexible polypeptides, di- and polyamines, and uncharged polymers. Thermal stability analysis of DNA oligonucleotide structures revealed that unlike conventional polymer cosolutes, basic globular proteins (lysozyme and cytochrome c) at high concentrations stabilized long internal and bulge loop structures but not fully matched duplexes. The selective stabilization of long loop structures suggests preferential binding to unpaired nucleotides in loops through weak electrostatic interactions. Furthermore, the ability of the proteins to stabilize the loop structures was enhanced under macromolecular crowding conditions. Remarkably, the effects of basic proteins on the stability of fully matched duplexes were dissimilar to those of basic amino-acid-rich polypeptides and polyamines. This study provides new insights into the interaction of nucleic acid structures with organic cations.  相似文献   

11.
A large set of protein structures resolved by X-ray or NMR techniques has been extracted from the Protein Data Bank and analyzed using statistical methods. In particular, we investigate the interactions between side chains and the interactions between solvent and side chains, pointing out on the possibility of including the solvent as part of a knowledge-based potential. The solvent-residue contacts are accounted for on the basis of the Voronoi's polyhedron analysis. Our investigation confirms the importance of hydrophobic residues in determining the protein stability. We observe that in general hydrophobic-hydrophobic interactions and, more specifically, aromatic-aromatic contacts tend to be increasingly distally separated in the primary sequence of proteins, thus connecting distinct secondary structure elements. A simple relation expressing the dependence of the protein free energy by the number of residues is proposed. Such a relation includes both the residue-residue and the solvent-residue contributions. The former is dominant for large size proteins, whereas for small sizes (number of residues less than 100) the two terms are comparable. Gapless threading experiments show that the solvent-residue knowledge-based potential yields a significant contribution with respect to discriminating the native structure of proteins. Such contribution is important especially for proteins of small size and is similar to that given by the most favorable residue-residue knowledge-based potential referring to hydrophobic-hydrophobic interactions such as isoleucine-leucine. In general, the inclusion of the solvent-residue interaction produces a relevant increase of the free energy gap between the native structures and decoys.  相似文献   

12.
The structural aspects of protein functions, e.g., molecular recognition such as enzyme-substrate and antibody-antigen interactions, are elucidated in terms of dehydration and atomic interactions. When a protein interacts with some target molecule, water molecules at the interacting regions of both molecules are removed, with loss of the hydration free energy, but gaining atomic interactions between atoms of the contact sites in both molecules. The free energies of association originating from the dehydration and interactions between the atoms can be computed from changes in the accessible surface areas of the atoms involved. The free energy due to interactions between atomic groups at the contact sites is estimated as the sum of those estimated from the changes in the accessible surface area of 7 atomic groups, assuming that the interactions are proportional to the change of the area. The chain enthalpies and entropies evaluated from experimental thermodynamic properties and hydration quantities at the standard temperature for 10 proteins were available to determine the proportional constants for the atomic groups. This method was applied to the evaluation of association constants for the dimerization of proteins and the formation of proteolytic enzyme-inhibitor complexes, and the computed constants were in agreement with the experimental ones. However, the method is not accurate enough to account quantitatively for the change in the thermal stability of mutants of T4 lysozyme. Nevertheless, this method provides a way to elucidate the interactions between molecules in solution.  相似文献   

13.
Synthetic and recombinant peptide models of the central repetitive domain of the high molecular weight subunits of wheat glutenin with different numbers of the consensus repeat motifs PGQGQQ + GYYPTSLQQ (21, 45, 110, and 203 residues long) and a recombinant 58,000-Da relative molecular mass (M(r) 58,000) repetitive peptide from a single subunit (1Dx5) are studied using Fourier transform IR spectroscopy. The spectra of the dry peptides are very similar; at low water contents (<76% relative humidity) there is an increase in beta-sheet structures in all peptides. However, on further hydration the content of beta sheets decrease and more beta turns are observed. The changes during the second step of hydration are very marked in the 21 and 45 residue peptides, but they are less apparent in the longer perfect repeat peptides. In the 110 and 203 residue peptides hydration results in increased contents of intermolecular beta-sheets and less beta-turn formation. In contrast, the beta-turn content of the M(r) 58,000 peptide increases during the second hydration step. The decreased extent of structure changes with increasing chain length indicates that cumulative intermolecular interactions, in particular hydrogen bonds, are an important factor in determining the structures in the solid state. The regularity of the perfect repeat sequences in the 21, 45, 110, and 203 residue peptides may favor the formation of larger stretches of intermolecular beta sheets. In contrast, the M(r) 58,000 peptide contains imperfect repeats (in common with native glutenin subunits), which may limit its ability to form intermolecular beta sheets.  相似文献   

14.
Vibrational Raman optical activity (ROA) spectra of the wheat proteins alpha-gliadin (A-gliadin), omega-gliadin, and a 30 kDa peptide called T-A-1 from the high molecular weight glutenin subunit (HMW-GS) Dx5 were measured to obtain new information about their solution structures. The spectral data show that, under the conditions investigated, A-gliadin contains a considerable amount of hydrated alpha-helix, most of which probably lies within a relatively structured C-terminal domain. Smaller quantities of beta-structure and poly(l-proline) II (PPII) helix were also identified. Addition of methanol was found to increase the alpha-helix content at the expense of some of the beta and PPII structure. In comparison, omega-gliadin and the T-A-1 peptide were found to consist of large amounts of well-defined PPII structure with some turns but no alpha-helix. The results for the T-A-1 peptide are in agreement with a model in which HMW-GS are extended but not highly rigid. Application of a pattern recognition technique, based on principal component analysis (PCA), to the ROA spectra reinforces these conclusions.  相似文献   

15.
The mutated hemoglobin HbC (beta 6 Glu-->Lys), in the oxygenated (R) liganded state, forms crystals inside red blood cells of patients with CC and SC diseases. Static and dynamic light scattering characterization of the interactions between the R-state (CO) HbC, HbA, and HbS molecules in low-ionic-strength solutions showed that electrostatics is unimportant and that the interactions are dominated by the specific binding of solutions' ions to the proteins. Microscopic observations and determinations of the nucleation statistics showed that the crystals of HbC nucleate and grow by the attachment of native molecules from the solution and that concurrent amorphous phases, spherulites, and microfibers are not building blocks for the crystal. Using a novel miniaturized light-scintillation technique, we quantified a strong retrograde solubility dependence on temperature. Thermodynamic analyses of HbC crystallization yielded a high positive enthalpy of 155 kJ mol(-1), i.e., the specific interactions favor HbC molecules in the solute state. Then, HbC crystallization is only possible because of the huge entropy gain of 610 J mol(-1) K(-1), likely stemming from the release of up to 10 water molecules per protein intermolecular contact-hydrophobic interaction. Thus, the higher crystallization propensity of R-state HbC is attributable to increased hydrophobicity resulting from the conformational changes that accompany the HbC beta 6 mutation.  相似文献   

16.
Lysobisphosphatidic acid (LBPA) can be regarded to represent a unique derivative of phosphatidylglycerol. This lipid is highly enriched in late endosomes where it can comprise up to 10-15 mol% of all lipids and in these membranes, LBPA appears to be segregated into microdomains. We studied the thermotropic behavior of pure dioleoyl-LBPA mono- and bilayers using Langmuir-lipid monolayers, electron microscopy, differential scanning calorimetry (DSC), and fluorescence spectroscopy. LBPA formed metastable, liquid-expanded monolayers at an air/buffer interface, and its compression isotherms lacked any indication for structural phase transitions. Neat LBPA formed multilamellar vesicles with no structural transitions or phase transitions between 10 and 80 degrees C at a pH range of 3.0-7.4. We then proceeded to study mixed LBPA/dipalmitoylphosphatidylcholine (DPPC) bilayers by DSC and fluorescence spectroscopy. Incorporating increasing amounts of LBPA (up to X(LBPA) (molar fraction)=0.10) decreased the co-operativity of the main transition for DPPC, and a decrease in the main phase transition as well as pretransition temperature of DPPC was observed yet with no effect on the enthalpy of this transition. In keeping with the DSC data for DPPC, 1-palmitoyl-2-oleoyl-phosphatidylcholine (POPC)/LBPA mixed bilayers were more fluid, and no evidence for lateral phase segregation was observed. These results were confirmed using fluorescence microscopy of Langmuir-lipid films composed of POPC and LBPA up to X(LBPA)=0.50 with no evidence for lateral phase separation. As late endosomes are eminently acidic, we examined the effect of lowering pH on lateral organization of mixed PC/LBPA bilayers by DSC and fluorescence spectroscopy. Even at pH 3.0, we find no evidence of LBPA-induced microdomain formation at LBPA contents found in cellular organelles.  相似文献   

17.
W. Zimmer  H. Bothe 《Plant and Soil》1988,110(2):239-247
A simple model system was designed to detect positive effects ofAzospirillum on the root growth of cereals. Cultures ofA. brasilense Sp7 andA. lipoferum Sp59 did not excrete gibberellins and cytokinins in the logarithmic and in the early stationary growth phase. Indoleacetic acid (IAA) was formed, however, only in the stationary phase of the cultures. The addition of D,L-tryptophan to the medium enhanced the formation of IAA. A further, still unidentified substance was produced byAzospirillum under denitrifying conditions in the logarithmic growth phase. The substance was almost twice as active as IAA in increasing the wet weight of wheat root segments. It is suggested that this unidentified substance is the major stimulus affecting the growth of cereals.Dedicated to Professor E.-G. Niemann, Hannover, on the occasion of his 60th birthday.  相似文献   

18.
The intermolecular interactions in concentrated solutions of pig submaxillary mucin (PSM) and sheep submaxillary mucin (SSM) were studied by mechanical spectroscopy. PSM and SSM were purified from detectable protein and nucleic acid by equilibrium centrifugation in a CsCl density gradient. PSM and SSM isolated in the presence of proteinase inhibitors showed distinct differences from preparations isolated in the presence of 0.2 M-NaCl alone, the latter having a carbohydrate and amino acid analysis similar to other preparations isolated by precipitation or ion-exchange techniques. Gel-filtration studies showed that preparations isolated in the presence of 0.2 M-NaCl alone were dissociated into smaller-sized glycoprotein units by 3.5 M-CsCl or 2.0 M-NaCl (SSM), pH 2.0 (PSM) or heating at 100 degrees C for 10 min (PSM and SSM). Preparations isolated in the presence of proteinase inhibitors were not dissociated by these treatments. Proteolysis fragmented all submaxillary mucin preparations into small glycopeptides of Mr 13,700 for PSM and of Mr 14,000 and 15,000 for SSM. PSM preparations when concentrated formed viscoelastic gels, as determined by mechanical spectroscopy. In contrast, SSM showed characteristics of a weak viscoelastic liquid under comparable conditions (coil overlap). PSM glycoprotein isolated in proteinase inhibitors formed weak viscoelastic gels at concentrations between 5 and 15 mg/ml. Preparations of PSM glycoprotein isolated in the presence of 0.2 M-NaCl (concentration 10-97 mg/ml) had the same overall mechanical gel structure as those preparations extracted in the presence of proteinase inhibitors. This gel structure was seen to collapse following proteolysis of both preparations or after acid treatment of the glycoprotein isolated in the presence of 0.2 M-NaCl, consistent with the breakdown in size of the polymeric glycoprotein. Treatment of PSM gel with 0.2 M-2-mercaptoethanol caused a surprising increase in gel strength, which was further markedly increased on removal of the reducing agent by dialysis. An association of reduced subunits of PSM was observed by gel filtration after removal of 0.2 M-2-mercaptoethanol. These results point to intermolecular disulphide exchange occurring on reduction of these PSM glycoprotein preparations. These results demonstrate that gel formation in PSM glycoprotein is similar to that for other gastrointestinal mucus glycoproteins from stomach to colon. Gel formation in PSM, as in other mucins, depends on polymerization of subunits.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
The interactions of fatty acids with proteins have been studied by a variety of conventional approaches for decades. However, only limited aspects of fatty acid-protein interactions have been elucidated, even with the integration of information gleaned from the many techniques. Judgments must be made about what information is most reliable, particularly when derivatives of fatty acids are substituted for natural fatty acids. In recent years, the application of techniques of structural biology has brought about dramatic advances in this important area of lipid research. High-resolution crystallographic and NMR structures of several proteins with bound fatty acids reveal the complete tertiary structure of the protein and molecular details of fatty acid-protein interactions. The examples presented include most of the known structures of (non-enzymatic) proteins that bind fatty acids. The proteins are found in very different compartments of cells and organisms: the plasma compartment (human serum albumin); the cytosolic compartment of mammalian cells (fatty acid- binding proteins); the cytosol of plant cells (nonspecific lipid-transfer protein); the nucleus of mammalian cells (peroxisome proliferator-activated receptor and hepatic nuclear factor 4); and a bacterial membrane (halorhodopsin). This review discusses the structural features of these proteins and their binding pocket(s) and compares the specific modes of their interactions with fatty acids.  相似文献   

20.
In order to understand the mechanisms behind the undesired aging of films based on vital wheat gluten plasticized with glycerol, films cast from water/ethanol solutions were investigated. The effect of pH was studied by casting from solutions at pH 4 and pH 11. The films were aged for 120 days at 50% relative humidity and 23 degrees C, and the tensile properties and oxygen and water vapor permeabilities were measured as a function of aging time. The changes in the protein structure were determined by infrared spectroscopy and size-exclusion and reverse-phase high-performance liquid chromatography, and the film structure was revealed by optical and scanning electron microscopy. The pH 11 film was mechanically more stable with time than the pH 4 film, the latter being initially very ductile but turning brittle toward the end of the aging period. The protein solubility and infrared spectroscopy measurements indicated that the protein structure of the pH 4 film was initially significantly less polymerized/aggregated than that of the pH 11 film. The polymerization of the pH 4 film increased during storage but it did not reach the degree of aggregation of the pH 11 film. Reverse-phase chromatography indicated that the pH 11 films were to some extent deamidated and that this increased with aging. At the same time a large fraction of the aged pH 11 film was unaffected by reducing agents, suggesting that a time-induced isopeptide cross-linking had occurred. This isopeptide formation did not, however, change the overall degree of aggregation and consequently the mechanical properties of the film. During aging, the pH 4 films lost more mass than the pH 11 films mainly due to migration of glycerol but also due to some loss of volatile mass. Scanning electron and optical microscopy showed that the pH 11 film was more uniform in thickness and that the film structure was more homogeneous than that of the pH 4 film. The oxygen permeability was also lower for the pH 11 film. The fact that the pH 4 film experienced a larger and more rapid change in its mechanical properties with time than the pH 11 film, as a consequence of a greater loss of plasticizer, was presumably due to its initial lower degree of protein aggregation/polymerization. Consequently, the cross-link density achieved at pH 4 was too low to effectively retain volatiles and glycerol within the matrix.  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