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 633 毫秒
1.
The crystal structures of recombinant and native beta homotrimers of soybean beta-conglycinin were determined by X-ray crystallography at 2.7 and 2.8 A resolutions, respectively. The crystals of the recombinant and native beta homotrimers belong to space group P21 with cell parameters a = 80.51 A, b = 63.48 A, c = 131.43 A, and beta = 90.01 degrees and with cell parameters a = 82.78 A, b = 69.47 A, c = 125.33 A and beta = 97.22 degrees, respectively. The beta monomers consist of amino-terminal and carboxyl-terminal modules that are very similar to each other and are related by a pseudo-dyad axis. Each module of the beta monomer is subdivided into a core and a loop domain. These structural features of both beta homotrimers are consistent with those of canavalin and phaseolin, which are similar vicilin class proteins. The superposition of the models of the native and recombinant beta monomers shows a root mean square deviation of 0.43-0.51 A for 343 common Calpha atoms within 2.0 A. This result indicates that the N-linked glycans do not influence the final structure of the beta homotrimer. Comparison of the models of beta-conglycinin, phaseolin and canavalin indicates that beta-conglycinin resembles canavalin rather than phaseolin, and that canavalin and phaseolin differ the most among them. The evolutional relationships are discussed.  相似文献   

2.
Park J  Kim J  Lee T  Lim M 《Biophysical journal》2008,94(11):L84-L86
Femtosecond vibrational spectroscopy was used to probe a functionally important dynamics and residual structure of myoglobin unfolded by 4 M guanidine HCl. The spectra of the dissociated CO indicated that the residual structure of unfolded myoglobin (Mb) forms a few hydrophobic cavities that could accommodate the dissociated ligand. Geminate rebinding (GR) of CO to the unfolded Mb is three-orders-of-magnitude faster and more efficient than the native Mb but similar to a model heme in a viscous solvent, suggesting that the GR of CO to heme is accelerated by the longer retention of the dissociated ligand near the Fe atom by the poorly-structured protein matrix of the unfolded Mb or viscous solvent. The inefficient GR of CO in native Mb, while dissociated CO is trapped in the primary heme pocket located near the active binding site, indicates that the tertiary structure of the pocket in native Mb plays a functionally significant role.  相似文献   

3.
The structure to 3.0 Å resolution of the protein canavalin from Jack beans has been determined by conventional X-ray crystallographic techniques. We present evidence that canavalin is a proteolytically modified form of the enzyme α-d-mannosidase. This is based on the facts that the purified precursor protein possesses substantial α-d-mannosidase activity; it comigrates on SDS-PAGE with Jack bean α-d-mannosidase prepared by other means; and its oligomer molecular weight and Zn2+ content are the same as those of Jack bean α-d-mannosidase.  相似文献   

4.
(—)-Epicatechin-3-gallate (ECG) and (— )-epigallocatechin-3-gallate (EGCG), major tea catechins, formed precipitates with soybean lipoxygenase (LOX) in the pH range of 4~7, although with accompanying 10 ~30% loss of the LOX activity. Yeast alcohol dehydrogenase also was precipitated by EGCG. Polyvinylpyrrolidone, Tween 20 and Triton X-100 dissociated the LOX activity from the EGCG-precipitated LOX. However, the MW of the dissociated LOX (114,000) differed from that of the native LOX (100,000). Enzyme activities of the EGCG-precipitated LOX and the dissociated LOX from the precipitate were less stable than the activity of the native LOX. These findings suggest the altered natures of proteins in the presence of tea catechins, ECG and EGCG.  相似文献   

5.
6.
Pig small-intestinal mucus glycoprotein, of molecular weight 1.72 X 10(6), is cleaved by Pronase digestion into glycoprotein subunits of molecular weight 4.5 X 10(5). Of the protein component of the native glycoprotein 29% by weight was lost on Pronase digestion, with no loss of carbohydrate. The non-glycosylated region of the protein that was lost with proteolytic digestion had a broad spectrum of amino acid residues, in contrast with the glycosylated region of the protein, which was resistant to proteolysis and was rich in serine, threonine and proline residues. Reduction with 0.2M-mercaptoethanol dissociated the Pronase-digested glycoprotein subunits into smaller glycoprotein subunits of molecular weight 2.7 X 10(5). On reduction, the native glycoprotein was dissociated into subunits of molecular weight 2.4 X 10(5), a similar size to those obtained from reduction of the Pronase-digested glycoprotein. On reductive dissociation of the native glycoprotein, in addition to glycoprotein subunits, protein was also released principally as a component of 90000 molecular weight. This protein was separated quantitatively from the reduced glycoprotein in amounts compatible with one 90000-mol.wt. protein molecule per 1.72 X 10(6)-mol.wt. native glycoprotein molecule. No 90000-mol.wt. protein was released on reduction of the isolated Pronase-digested glycoprotein. Pig small-intestinal mucus glycoprotein is therefore a covalent polymer of glycoprotein subunits joined by disulphide bridges. This polymeric structure differs in important respects from that previously shown for gastric mucus, in particular with respect to the size and number of component subunits per native molecule.  相似文献   

7.
In situ laser Michelson interferometry was utilized to investigate mechanisms of growth and surface morphology in protein and virus crystallization, These included plant proteins canavalin and thaumatin and turnip yellow mosaic virus. The experimental apparatus allowed us to obtain interferometric patterns and investigate growth kinetics from growing macromolecular crystals as small as 20 μm. For the crystallization of canavalin, dislocations are the sources of growth steps on the surfaces. Supersaturation and time dependencies of the normal growth rates, tangential growth step velocities, and the slopes of the dislocation hillocks were measured. The kinetic coefficient β (rate of incorporation of protein molecules into the growing crystal) was estimated for canavalin to be 9 × 10-4 cm/sec. This is among the first estimates of such fundamental kinetic parameters for macromolecular crystallization. The change in the activities of dislocation sources under different growth conditions was also analyzed. Michelson interferometry was clearly demonstrated to be a useful tool for quantitative studies of macromolecular crystal growth.  相似文献   

8.
Methylation in vivo is a post-translational modification observed in several organisms belonging to eucarya, bacteria, and archaea. Although important implications of this modification have been demonstrated in several eucaryotes, its biological role in hyperthermophilic archaea is far from being understood. The aim of this work is to clarify some effects of methylation on the properties of beta-glycosidase from Sulfolobus solfataricus, by a structural comparison between the native, methylated protein and its unmethylated counterpart, recombinantly expressed in Escherichia coli. Analysis by Fourier transform infrared spectroscopy indicated similar secondary structure contents for the two forms of the protein. However, the study of temperature perturbation by Fourier transform infrared spectroscopy and turbidimetry evidenced denaturation and aggregation events more pronounced in recombinant than in native beta-glycosidase. Red Nile fluorescence analysis revealed significant differences of surface hydrophobicity between the two forms of the protein. Unlike the native enzyme, which dissociated into SDS-resistant dimers upon exposure to the detergent, the recombinant enzyme partially dissociated into monomers. By electrospray mapping, the methylation sites of the native protein were identified. A computational analysis of beta-glycosidase three-dimensional structure and comparisons with other proteins from S. solfataricus revealed analogies in the localization of methylation sites in terms of secondary structural elements and overall topology. These observations suggest a role for the methylation of lysyl residues, located in selected domains, in the thermal stabilization of beta-glycosidase from S. solfataricus.  相似文献   

9.
The number of artificial protein supramolecules has been increasing; however, control of protein oligomer formation remains challenging. Cytochrome c′ from Allochromatium vinosum (AVCP) is a homodimeric protein in its native form, where its protomer exhibits a four‐helix bundle structure containing a covalently bound five‐coordinate heme as a gas binding site. AVCP exhibits a unique reversible dimer–monomer transition according to the absence and presence of CO. Herein, domain‐swapped dimeric AVCP was constructed and utilized to form a tetramer and high‐order oligomers. The X‐ray crystal structure of oxidized tetrameric AVCP consisted of two monomer subunits and one domain‐swapped dimer subunit, which exchanged the region containing helices αA and αB between protomers. The active site structures of the domain‐swapped dimer subunit and monomer subunits in the tetramer were similar to those of the monomer subunits in the native dimer. The subunit–subunit interactions at the interfaces of the domain‐swapped dimer and monomer subunits in the tetramer were also similar to the subunit–subunit interaction in the native dimer. Reduced tetrameric AVCP dissociated to a domain‐swapped dimer and two monomers upon CO binding. Without monomers, the domain‐swapped dimers formed tetramers, hexamers, and higher‐order oligomers in the absence of CO, whereas the oligomers dissociated to domain‐swapped dimers in the presence of CO, demonstrating that the domain‐swapped dimer maintains the CO‐induced subunit dissociation behavior of native ACVP. These results suggest that protein oligomer formation may be controlled by utilizing domain swapping for a dimer–monomer transition protein.  相似文献   

10.
The structural bovine eye lens protein alpha-crystallin was dissociated in 7 M urea and its four subunits, A1, A2, B1, and B2, were separated by means of ion-exchange chromatography. Homopolymeric reaggregates of these subunits were prepared by removal of the denaturant via dialysis. It was found that subunits were exchanged upon incubation of mixtures of two homopolymers under native conditions. New hybrid species were formed within 24 h as demonstrated by isoelectric focusing. Moreover, native alpha-crystallin molecules also exchanged subunits when incubated with homopolymeric aggregates of B2 subunits. Subunit exchange between native alpha-crystallin molecules is postulated, and a "dynamic quaternary structure" is presented that allows the polydisperse protein to adapt to changes in cytoplasmic conditions upon aging of the lens tissue.  相似文献   

11.
Canavalin is a vicilin-class (7S) storage protein found in sword bean (Canavalia gladiata). Our previous report indicated that canavalin is precipitated by the addition of 20 mM MgCl2 to crude sword bean extract. Here, we examined the solubility changes induced by the addition of Mg2+ and Ca2+ at various concentrations. Canavalin tended to be insolubilized at relatively low concentrations of MgCl2 (< 20 mM) and solubilized at relatively high concentrations (> 20 mM). In addition, canavalin was slightly insolubilized in the presence of NaCl. Overall, the results revealed that solubility changes are reversible and depend on the concentration of divalent cations. Therefore, we suggested a reaction scheme that describes the effects of divalent cations on the solubility of canavalin, which would facilitate the study of its physiological function and the application of canavalin in the food processing industry.  相似文献   

12.
Wild-type and deglycosylated forms of human prostate-specific antigen were expressed in Chinese hamster ovary (CHO) cells as zymogens. ProPSA was collected from conditioned medium and purified using a single cation-exchange chromatographic step for the deglycosylated form and cation-exchange followed by gel filtration chromatography for the wild-type form. Recombinant wild-type proPSA produced in CHO cells has an average MW of 34.5 kDa, whereas the deglycosylated proPSA has a MW of 32.4 kDa. Both forms of proPSA were activated in vitro and the kinetic properties measured for the deglycosylated PSA are very similar to those of the wild-type recombinant PSA and the native PSA isolated from seminal fluid. These results suggest that deglycosylated PSA is likely to be very similar to native PSA with respect to its three-dimensional structure and will provide a homogeneous protein preparation necessary for X-ray crystallographic analysis.  相似文献   

13.
The GroES protein from Escherichia coli is a well-known member of the molecular chaperones. GroES consists of seven identical 10 kDa subunits, and forms a dome-like oligomeric structure. In order to obtain information on the structural stability and unfolding-refolding mechanism of GroES protein, especially at protein concentrations (0.4-1.2 mM GroES monomer) that would mimic heat stress conditions in vivo, we have performed synchrotron small-angle X-ray scattering (SAXS) experiments. Surprisingly, in spite of the high protein concentration, reversibility in the unfolding-refolding reaction was confirmed by SAXS experiments structurally. Although the unfolding-refolding reaction showed an apparent single transition with a Cm of 1.1 M guanidium hydrochloride, a more detailed analysis of this transition demonstrated that the unfolding mechanism could be best explained by a sequential three-state model, which consists of native heptamer, dissociated monomer, and unfolded monomer. Together with our previous result that GroES unfolded completely via a partially folded monomer according to a three-state model at low protein concentration (5 microM monomer), the unfolding-refolding mechanism of GroES protein could be explained uniformly by the three-state model from low to high protein concentrations. Furthermore, to clarify an ambiguity of the native GroES structure in solution, especially mobile loop structures, we have estimated a solution structure of GroES using SAXS profiles obtained from experiments and simulation analysis. The result suggested that the native structure of GroES in solution was very similar to that seen in GroES-GroEL complex determined by crystallography.  相似文献   

14.
The aggregation processes leading to crystallization and precipitation of canavalin have been investigated by dynamic light scattering (DLS) in photon correlation spectroscopy (PCS) mode. The sizes of aggregates formed under various conditions of pH, salt concentration, and protein concentrations were deduced from the correlation functions generated by the fluctuating intensity of light scattered by the solutions of the protein. Results obtained indicate that the barrier to crystallization of canavalin is the formation of the trimer, a species that has been characterized by x-ray crystallographic studies (McPherson, A. 1980. J. Biol. Chem. 255:10472-10480). The dimensions of the trimer in solution are in good agreement with those obtained both from the crystal (McPherson, A. 1980. J. Biol. Chem. 255:10472-10480) and from a low angle x-ray scattering study in solution (Plietz, P., P. Damaschun, J. J. Müller, and B. Schlener. 1983. FEBS [Fed. Eur. Biochem. Soc.] Lett. 162:43-46). Furthermore, under conditions known to lead to the formation of rhombohedral crystals of canavalin, a limiting size is reached at high concentrations of canavalin. The size measured corresponds to an aggregate of trimers making a unit rhombohedral cell consistent with x-ray crystallographic data (McPherson, A. 1980. J. Biol. Chem. 255:10472-10480). Presumably, such aggregates are the nuclei from which crystal growth proceeds. The present study was undertaken primarily to test the potential of DLS (PCS) as a tool for rapid, routine screening to determine the ultimate fate of protein solutions (i.e., crystallization or amorphous precipitation) at an early stage, therefore eliminating the need for long-term visual observation.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
We determined the three-dimensional crystal structure of the protein YML079wp, encoded by a hypothetical open reading frame from Saccharomyces cerevisiae to a resolution of 1.75 A. The protein has no close homologs and its molecular and cellular functions are unknown. The structure of the protein is a jelly-roll fold consisting of ten beta-strands organized in two parallel packed beta-sheets. The protein has strong structural resemblance to the plant storage and ligand binding proteins (canavalin, glycinin, auxin binding protein) but also to some plant and bacterial enzymes (epimerase, germin). The protein forms homodimers in the crystal, confirming measurements of its molecular mass in solution. Two monomers have their beta-sheet packed together to form the dimer. The presence of a hydrophobic ligand in a well conserved pocket inside the barrel and local sequence similarity with bacterial epimerases may suggest a biochemical function for this protein.  相似文献   

16.
Sword bean (Canavalia gladiata) seeds are a traditional food in Asian countries. In this study, we aimed to determine the optimal methods for the precipitation of sword bean proteins useful for the food development. The soaking time for sword beans was determined by comparing it with that for soybeans. Sword bean proteins were extracted from dried seeds in distilled water using novel methods. We found that most proteins could be precipitated by heating the extract at more than 90 °C. Interestingly, adding magnesium chloride to the extract at lower temperatures induced specific precipitation of a single protein with a molecular weight of approximately 48 kDa. The molecular weight and N-terminal sequence of the precipitated protein was identical to that of canavalin. These data suggested that canavalin was precipitated by the addition of magnesium chloride to the extract. Our results provide important insights into the production of processed foods from sword bean.  相似文献   

17.
The structure of canavalin, a jack bean (Canavalis ensiformis) protein homologous to phaseolin, the major seed storage protein of Phaseolus vulgaris, has been investigated by x-ray crystallography and found to be a hexamer composed of three identical pairs of similar but nonidentical subunits related by a perfect 3-fold axis and pseudo dyad axes (strict C3 and pseudo D3). One member of each pair of subunits is derived from the amino terminal half of a precursor polypeptide of molecular weight 49,000 and the other from its carboxy terminal half. Thus, the crystallographic evidence indicates that the precursor polypeptide is a tandem duplicate and is structurally redundant (McPherson A. 1982 J Biol Chem 255: 10472). A number of physical and chemical properties of the protein in both the uncleaved and the cleaved form were investigated. These included the native molecular weights, amino acid analyses, number of exposed sulfhydryl groups, carbohydrate content, metal ion analysis, crystallization behavior, and the fate of the protein in developing seeds. It was also found that the purified precursor protein possesses a substantial level of α-d-mannosidase activity and seems to share a number of other physical and chemical properties with that enzyme.  相似文献   

18.
We examined the primary sequence of canavalin, the major storage protein of jack beans, and found that an ancient sequence duplication accounts for 80% of the amino acid residues. Evidence for such a duplication was also found in the orthologous proteins phaseolin and pea vicilin. This sequence duplication presumably accounts for a structural duplication in the canavalin monomer observed by crystallographic analysis. One copy of this repeat was found in a second storage-protein family, the legumins, where it encompasses almost the entire B-chain of the mature molecule. We propose that the vicilin and legumin families of legume seed proteins evolved from a common precursor, which consisted of one copy of the repeat in the vicilins.  相似文献   

19.
T P Ko  J D Ng    A McPherson 《Plant physiology》1993,101(3):729-744
The three-dimensional structure of the vicilin storage protein canavalin, from Canavalia ensiformis, has been determined in a hexagonal crystal by x-ray diffraction methods. The model has been refined at 2.6 A resolution to an R factor of 0.197 with acceptable geometry. Because of proteolysis, 58 of 419 amino acids of the canavalin polypeptide are not visible in the electron density map. The canavalin subunit is composed of two extremely similar structural domains that reflect the tandem duplication observed in the cDNA and in the amino acid sequence. Each domain consists of two elements, a compact, eight-stranded beta-barrel having the "Swiss roll" topology and an extended loop containing several short alpha-helices. The root mean square deviation between 84 pairs of corresponding C alpha atoms making up the strands of the two beta-barrels in a subunit is 0.78 A, and for 112 pairs of structurally equivalent C alpha atoms of the two domains the deviation is 1.37 A. The interface between domains arises from the apposition of broad hydrophobic surfaces formed by side chains originating from one side of the beta-barrels, supplemented by at least four salt bridges. The interfaces between subunits in the trimer are supplied by the extended loop elements. These interfaces are also composed primarily of hydrophobic residues supplemented by six salt bridges. The canavalin subunits have dimensions about 40 x 40 x 86 A, and the oligomer is a disk-shaped molecule about 88 A in diameter with a thickness of about 40 A. The distribution of domains lends a high degree of pseudo-32-point group symmetry to the molecule. There is a large channel of 18 A diameter, lined predominantly by hydrophilic and charged amino acids, running through the molecule along the 3-fold axis. The majority of residues conserved between domains and among vicilins occur at the interface between subunits but appear otherwise arbitrarily distributed within the subunit, although predominantly on its exterior.  相似文献   

20.
The regulatory subunit of cAMP-dependent protein kinase has a well-defined domain structure, and recombinant DNA techniques have been used to define further the functional properties that are associated with each domain. Our initial question was to define the minimal structural unit that is required for forming a stable complex with the catalytic subunit that will still bind and hence be dissociated by cAMP. To answer these questions, the entire second cAMP-binding domain was deleted using oligonucleotide-directed mutagenesis to introduce a premature stop codon at Trp260. This mutation results in the expression of a stable protein with an Mr of 38,000 based on polyacrylamide gel electrophoresis. The resulting mutant protein is a dimer; and like the native R-subunit, the two protomers of the dimer are cross-linked by disulfide bonds at the amino terminus. The mutant R-subunit binds 1 mol of cAMP/monomer based on equilibrium dialysis. The Kd(cAMP) was 25 nM, which is slightly higher than the Kd(cAMP) for the native R-subunit. The removal of the second cAMP domain does not prevent aggregation with the catalytic subunit, and the inactive holoenzyme complex that is formed in the absence of cAMP can still be dissociated and consequently activated by cAMP. In conjunction with previous results based on limited proteolysis, it is concluded that the region extending from Arg94 to Lys259 constitutes a structural unit that will be sufficient to interact with the catalytic subunit in a cAMP-dependent manner.  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