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.
Abstract The nucleotide sequence of the Rhodobacter capsulatus bacterioferritin gene ( bfr ) was determined and found to encode a protein of 161 amino acids with a predicted molecular mass of 18 174 Da. The molecular mass of the purified protein was estimated to be 18 176.06 ± 0.80 Da by electrospray mass spectrometry. The bfr gene was introduced into an expression vector, and bacterioferritin was produced to a high level in Escherichia coli . The amino acids which are involved in haem ligation, and those which provide ligands in the binuclear metal centre in bacterioferritin from E. coli are conserved in the R. capsulatus protein. The sequences of bacterioferritins, ferritin-like proteins, and proteins similar to Dps of E. coli are compared, and membership of the bacterioferritin family re-evaluated.  相似文献   

2.
Urea-induced dissociation and unfolding of manganese.glutamine synthetase (Mn.GS) have been studied at 37 degrees C (pH 7) by spectroscopic and calorimetric methods. In 0 to approximately 2 M urea, Mn.GS retains its dodecameric structure and full catalytic activity. Mn.GS is dissociated into subunits in 6 M urea, as evidenced by a 12-fold decrease in 90 degrees light scattering and a monomer molecular weight of 51,800 in sedimentation equilibrium studies. The light scattering decrease in 4 M urea parallels the time course of Trp exposure but occurs more rapidly than changes in secondary structure and Tyr exposure. Early and late kinetic steps appear to involve predominantly disruption of intra-ring and inter-ring subunit contacts, respectively, in the layered hexagonal structure of Mn.GS. The enthalpies for transferring Mn.GS into urea solutions have been measured by titration calorimetry. After correcting for the enthalpy of binding urea to the protein, the enthalpy of dissociation and unfolding of Mn.GS is 14 +/- 4 cal/g. A net proton uptake of approximately 50 H+/dodecamer accompanies unfolding reactions. The calorimetric data are consistent with urea binding to multiple, independent sites in Mn.GS and the number of binding sites increasing approximately 9-fold during the protein unfolding.  相似文献   

3.
Cold denaturation is a phenomenon seen in many different proteins. However, there have been no reports so far of its occurrence in hyperthermophile proteins. Here, using a recombinant triosephosphate isomerase (PfuTIM) from the hyperthermophile archaeon, Pyrococcus furiosus, we show that the heating of this protein through the low temperature side of its thermal unfolding transition in the presence of guanidinium hydrochloride (GdmCl) results in the formation of partially-disordered conformational ensembles that retain considerable native-like secondary and tertiary structure. Unlike PfuTIM itself, these thermochemically obtained partially-disordered PfuTIM ensembles display cold denaturation as they are cooled to room temperature. The protein thus shows hysteresis, adopting different structural states in a manner dependent upon the nature of the heating and cooling treatment, rather than upon the initial and final conditions of temperature and GdmCl concentration, indicating that some sort of a kinetic effect influences structure adoption and retention. The structure lost through cooling of partially-disordered PfuTIM is found to be regained through heating. The ability of GdmCl to thus apparently destabilize the highly thermodynamically and kinetically stable structure of PfuTIM (sufficiently, to cause it to display observable cold-denaturation and heat-renaturation transitions, in real-time, with cooling and heating) offers support to current ideas concerning the how hyperthermophile proteins achieve their high kinetic stabilities, and suggests that desolvation-solvation barriers may be responsible for high kinetic stability.  相似文献   

4.
Nonspecific lipid transfer protein (nsLTP; also called sterol carrier protein 2) with a molecular mass of 13 kDa is synthesized as a larger 15-kDa precursor (pre-nsLTP) with an N-terminal 20-amino acid extension presequence, as well as with the peroxisome targeting signal type 1 (PTS1), Ala-Lys-Leu, at the C terminus. The precursor pre-nsLTP is processed to mature nsLTP by proteolytic removal of the presequence, most likely after being imported into peroxisomes. Sterol carrier protein x (SCPx), a 59-kDa branched-chain fatty acid thiolase of peroxisomes, contains the entire pre-nsLTP moiety at the C-terminal part and is converted to the 46-kDa form and nsLTP after the transport to peroxisomes. We investigated which of these two potential topogenic sequences functions in biogenesis of nsLTP and SCPx. Morphological and biochemical analyses, making use of Chinese hamster ovary cell pex mutants such as the PTS1 receptor-impaired pex5 and PTS2 import-defective pex7, as well as green fluorescent protein chimeras, revealed that both pre-nsLTP and SCPx are imported into peroxisomes by the Pex5p-mediated PTS1 pathway. Nearly half of the pre-nsLTP remains in the cytosol, as assessed by subcellular fractionation of the wild-type Chinese hamster ovary cells. In an in vitro binding assay, only mature nsLTP, but not pre-nsLTP, from the cell lysates interacted with the Pex5p. It is likely, therefore, that modulation of the C-terminal PTS1 by the presequence gives rise to cytoplasmic localization of pre-nsLTP.  相似文献   

5.
6.
Gibberellins A4, A9, and A24 were isolated and identified from the new gibberellin-producing fungus, Phaeosphaeria sp. L487 (Loculoascomycetes), cultivated in a chemically defined medium; their yields from the culture filtrate were ca. 1. 7, 0. 3, and 0.4 μg/ml, respectively. Gibberellins A4 and A9 significantly stimulated the hypocotyl growth of Chinese cabbage seedlings at a very low concentration of less than 0.01 μg/ml.  相似文献   

7.
8.
Iwaki M  Osyczka A  Moser CC  Dutton PL  Rich PR 《Biochemistry》2004,43(29):9477-9486
Redox transitions in the Rhodobacter capsulatus cytochrome bc(1) complex were investigated by perfusion-induced attenuated total reflection-Fourier transform infrared (ATR-FTIR) spectroscopy combined with synchronous visible spectroscopy, in both the wild type and a cytochrome c(1) point mutant, M183K, in which the midpoint potential of heme was lowered from the wild-type value of 320 mV to 60 mV. Overall redox difference spectra of the wild type and M183K mutant were essentially identical, indicating that the mutation did not cause any major structural perturbation. Spectra were compared with data on the bovine bc(1) complex, and tentative assignments of several bands could be made by comparison with available data on model compounds and crystallographic structures. The bacterial spectra showed contributions from ubiquinone that were much larger than in the bovine enzyme, arising from additional bound and adventitious ubiquinone. The M183K mutant enabled selective reduction of the iron-sulfur protein which in turn allowed the IR redox difference spectra of ISP and cytochrome c(1) to be deconvoluted at high signal/noise ratios, and features of these spectra are interpreted in light of structural and mechanistic information.  相似文献   

9.
α‐Crystallin is a multimeric eye lens protein having molecular chaperone‐like function which is crucial for lens transparency. The stability and unfolding‐refolding properties of α‐crystallin plays important roles for its function. We undertook a multi probe based fluorescence spectroscopic approach to explore the changes in the various levels of organization of this protein at different urea concentration. Steady state fluorescence studies reveal that at 0.6M urea a compact structural intermediate is formed which has a native‐like secondary structure with enhanced surface exposure of hydrophobic groups. At 2.8M urea the tertiary interactions are largely collapsed with partial collapse of secondary and quaternary structure. The surface solvation probed by picosecond time resolved fluorescence of acrylodan labeled α‐crystallin revealed dry native‐like core of α‐crystallin at 0.6M urea compared to enhanced water penetration at 2.8M urea and extensive solvation at 6M urea. Activation energy for the subunit exchange decreased by 22 kJ mol?1 on changing urea concentration from 0 to 0.6M compared with over 75 kJ mol?1 on changing urea concentration from 0 to 2.8M. Light scattering and analytical ultracentrifugation techniques were used to determine size and oligomerization of the unfolding intermediates. The data indicated swelling but no oligomer breakdown at 0.6M urea. At 2.8M urea the oligomeric size is considerably reduced and a monomer is produced at 6M urea. The data clearly reveals that structural breakdown of α‐crystallin does not follow hierarchical sequence as tertiary structure dissolution takes place before complete oligomeric dissociation. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 549–560, 2014.  相似文献   

10.
11.
Campbell JC  Whitten ST 《Proteins》2012,80(1):184-193
Characterizing the native ensemble of protein is an important yet difficult objective of structural biology. The structural dynamics of protein macromolecules play key roles in biological function, but the short lifetimes and low population of near-native states of the protein ensemble limit their ability to be studied directly. In part to address such issues, it was shown recently that the cooperative substructures that populate a protein ensemble could be ascertained by NMR methods performed at very cold temperatures. What is presented here is an argument that these same substructures can also be determined by denaturant-induced unfolding studies performed on protein at room temperature. Data supporting this argument are given for Staphylococcal nuclease, chymotrypsin inhibitor 2, and ubiquitin. The observation of an agreement between the thermodynamics of the protein ensemble simulated under very cold temperatures to the apparent sensitivity of the ensemble to chemical denaturants at room temperature also suggests that the overall structural-thermodynamic character of an ensemble is surprisingly robust and preserved even in the presence of strong denaturing conditions.  相似文献   

12.
Recently, we presented a convenient method combining a deuterium‐hydrogen exchange and electrospray mass spectrometry for studying high‐pressure denaturation of proteins (Stefanowicz et al., Biosci Rep 2009; 30:91–99). Here, we present results of pressure‐induced denaturation studies of an amyloidogenic protein—the wild‐type human cystatin C (hCC) and its single‐point mutants, in which Val57 residue from the hinge region was substituted by Asn, Asp or Pro, respectively. The place of mutation and the substituting residues were chosen mainly on a basis of theoretical calculations. Observation of H/D isotopic exchange proceeding during pressure induced unfolding and subsequent refolding allowed us to detect differences in the proteins stability and folding dynamics. On the basis of the obtained results we can conclude that proline residue at the hinge region makes cystatin C structure more flexible and dynamic, what probably facilitates the dimerization process of this hCC variant. Polar asparagine does not influence stability of hCC conformation significantly, whereas charged aspartic acid in 57 position makes the protein structure slightly more prone to unfolding. Our experiments also point out pressure denaturation as a valuable supplementary method in denaturation studies of mutated proteins. Proteins 2012;. © 2012 Wiley Periodicals, Inc.  相似文献   

13.
With advancements in high‐throughput generation of phenotypic data on mutant proteins, it has become important to individually characterize different proteins or their variants rapidly and with minimal sample consumption. We have made use of a nano differential scanning fluorimetric device, from NanoTemper technologies, to rapidly carry out isothermal chemical denaturation and measure folding/unfolding kinetics of proteins and compared these to corresponding data obtained from conventional spectrofluorimetry. We show that using sample volumes 10‐50‐fold lower than with conventional fluorimetric techniques, one can rapidly and accurately measure thermodynamic and kinetic stability, as well as folding/unfolding kinetics. This method also facilitates characterization of proteins that are difficult to express and purify.  相似文献   

14.
A fundamental understanding of protein stability and the mechanism of denaturant action must ultimately rest on detailed knowledge about the structure, solvation, and energetics of the denatured state. Here, we use (17)O and (2)H magnetic relaxation dispersion (MRD) to study urea-induced denaturation of intestinal fatty acid-binding protein (I-FABP). MRD is among the few methods that can provide molecular-level information about protein solvation in native as well as denatured states, and it is used here to simultaneously monitor the interactions of urea and water with the unfolding protein. Whereas CD shows an apparently two-state transition, MRD reveals a more complex process involving at least two intermediates. At least one water molecule binds persistently (with residence time >10 nsec) to the protein even in 7.5 M urea, where the large internal binding cavity is disrupted and CD indicates a fully denatured protein. This may be the water molecule buried near the small hydrophobic folding core at the D-E turn in the native protein. The MRD data also provide insights about transient (residence time <1 nsec) interactions of urea and water with the native and denatured protein. In the denatured state, both water and urea rotation is much more retarded than for a fully solvated polypeptide. The MRD results support a picture of the denatured state where solvent penetrates relatively compact clusters of polypeptide segments.  相似文献   

15.
A glutathione S-transferase (Sj26GST) from Schistosoma japonicum, which functions in the parasite's Phase II detoxification pathway, is expressed by the Pharmacia pGEX-2T plasmid and is used widely as a fusion-protein affinity tag. It contains all 217 residues of Sj26GST and an additional 9-residue peptide linker with a thrombin cleavage site at its C-terminus. Size-exclusion HPLC (SEC-HPLC) and SDS-PAGE studies indicate that purification of the homodimeric protein under nonreducing conditions results in the reversible formation of significant amounts of 160-kDa and larger aggregates without a loss in catalytic activity. The basis for oxidative aggregation can be ascribed to the high degree of exposure of the four cysteine residues per subunit. The conformational stability of the dimeric protein was studied by urea- and temperature-induced unfolding techniques. Fluorescence-spectroscopy, SEC-HPLC, urea- and temperature-gradient gel electrophoresis, differential scanning microcalorimetry, and enzyme activity were employed to monitor structural and functional changes. The unfolding data indicate the absence of thermodynamically stable intermediates and that the unfolding/refolding transition is a two-state process involving folded native dimer and unfolded monomer. The stability of the protein was found to be dependent on its concentration, with a delta G degree (H2O) = 26.0 +/- 1.7 kcal/mol. The strong relationship observed between the m-value and the size of the protein indicates that the amount of protein surface area exposed to solvent upon unfolding is the major structural determinant for the dependence of the protein's free energy of unfolding on urea concentration. Thermograms obtained by differential scanning microcalorimetry also fitted a two-state unfolding transition model with values of delta Cp = 7,440 J/mol per K, delta H = 950.4 kJ/mol, and delta S = 1,484 J/mol.  相似文献   

16.
The effects of chain cleavage and circular permutation on the structure, stability, and activity of dihydrofolate reductase (DHFR) from Escherichia coli were investigated by various spectroscopic and biochemical methods. Cleavage of the backbone after position 86 resulted in two fragments, (1--86) and (87--159) each of which are poorly structured and enzymatically inactive. When combined in a 1 : 1 molar ratio, however, the fragments formed a high-affinity (K(a) = 2.6 x 10(7) M(-1)) complex that displays a weakly cooperative urea-induced unfolding transition at micromolar concentrations. The retention of about 15% of the enzymatic activity of full-length DHFR is surprising, considering that the secondary structure in the complex is substantially reduced from its wild-type counterpart. In contrast, a circularly permuted form with its N-terminus at position 86 has similar overall stability to full-length DHFR, about 50% of its activity, substantial secondary structure, altered side-chain packing in the adenosine binding domain, and unfolds via an equilibrium intermediate not observed in the wild-type protein. After addition of ligand or the tight-binding inhibitor methotrexate, both the fragment complex and the circular permutant adopt more native-like secondary and tertiary structures. These results show that changes in the backbone connectivity can produce alternatively folded forms and highlight the importance of protein-ligand interactions in stabilizing the active site architecture of DHFR.  相似文献   

17.
A pufX gene deletion in the purple bacterium Rhodobacter capsulatus causes a severe photosynthetic defect and increases core light-harvesting complex (LH1) protein and bacteriochlorophyll a (BChl) levels. It was suggested that PufX interrupts the LH1 alpha/beta ring around the reaction centre, allowing quinone/quinol exchange. However, naturally PufX(-) purple bacteria grow photosynthetically with an uninterrupted LH1. We discovered that substitutions of the Rhodobacter-specific LH1 alpha seryl-2 decrease carotenoid levels in PufX(-)R. capsulatus. An LH1 alphaS2F mutation improved the photosynthetic growth of a PufX(-) strain lacking the peripheral LH2 antenna, although LH1 BChl absorption remained above wild-type, suggesting that Rhodobacter-specific carotenoid binding is involved in the PufX(-) photosynthetic defect and LH1 expansion is not. Furthermore, PufX overexpression increased LH1-like BChl absorption without inhibiting photosynthetic growth. PufX(+) LH1 alphaS2-substituted mutant strains had wild-type carotenoid levels, indicating that PufX modulates LH1 carotenoid binding, inducing a conformational change that favours quinone/quinol exchange.  相似文献   

18.
Piscidin 1 (Pis‐1) has a high broad‐spectrum activity against bacteria, fungi, and viruses but it also has a moderate hemolytic activities. To improve the antibacterial activity and to reduce toxicity, mutants Pis‐1AA (G8A/G13A double mutant) and Pis‐1PG (G8P mutant) have been designed based on the crystal structure of Pis‐1. Eighteen independent molecular dynamics (MD) simulations of Pis‐1 and its mutants with membranes are conducted in this article. Furthermore, 60 independent MD simulations of three peptides in water box have also been discussed for comparison. The results indicate that the unfolding process starts at the middle of the peptide. Pis‐1 disrupts easily in the region of Val10‐Lys14. Pis‐1PG has a flexible N‐terminal region, and the interaction between N‐terminal and C‐terminal is very weak. Pis‐1AA has the most stable helical structure. In addition, percentage of native contacts and hydrogen bonds analysis are also performed. Lipid‐peptide interaction analysis suggests that Pis‐1 and Pis‐1AA has a stronger interaction with the zwitterionic dioleoylphosphatidylcholine (DOPC) lipid bilayer than Pis‐1PG. When compared with the results of peptide with membrane, peptides are unstable and unfolding quickly in water solution. Our results are applicable in examining diversities on hemolytic, antibacterial, and selectivity of antimicrobial peptides. © 2012 Wiley Periodicals, Inc. Biopolymers 97:998–1009, 2012.  相似文献   

19.
A mature mutant ribose-binding protein (RBP) ofEscherichia coli was obtained by site-directed mutagenesis, replacing Thr-3 in the N-domain of wild-type mature RBP (WT-mRBP) with a Trp residue (N-Trp-mRBP). The equilibrium unfolding properties and the refolding kinetics of this protein were monitored by fluorescence and circular dichroism (CD). The stability of N-Trp-mRBP appears to be the same as that of C-Trp-mRBP, another mutant obtained by replacing Phe-187 with a Trp, and lower than that of WT-mRBP. The overall refolding rate of N-Trp-mRBP is much smaller than that of C-Trp-mRBP, which, in turn, is similar to that of WT-mRBP. For the case of WT-mRBP, the rate constant obtained by Tyr fluorescence is identical to the value obtained by CD. But with C-Trp-mRBP, the rate constant from CD is smaller than the value from the Trp fluorescence and this difference in the rate constants is much greater with the N-TrpmRBP.  相似文献   

20.
Thermal transitions of many proteins have been found to be calorimetrically irreversible and scan-rate dependent. Calorimetric determinations of stability parameters of proteins which unfold irreversibly according to a first-order kinetic scheme have been reported. These methods require the approximation that the increase in heat capacity upon denaturation deltaCp is zero. A method to obtain thermodynamic parameters and activation energy for the two-state irreversible process N --> D from nonlinear fitting to calorimetric traces is proposed here. It is based on a molar excess heat capacity function which considers irreversibility and a nonzero constant deltaCp. This function has four parameters: (1) temperature at which the calorimetric profile reaches its maximal value (Tm), (2) calorimetric enthalpy at Tm (deltaHm), (3) deltaCp, and (4) activation energy (E). The thermal irreversible denaturation of subtilisin BPN' from Bacillus amyloliquefaciens was studied by differential scanning calorimetry at pH 7.5 to test our model. Transitions were found to be strongly scanning-rate dependent with a mean deltaCp value of 5.7 kcal K(-1)mol(-1), in agreement with values estimated by accessible surface area and significantly higher than a previously reported value.  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