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 31 毫秒
1.
Battistutta R  Negro A  Zanotti G 《Proteins》2000,41(4):429-437
The mutant F99S/M153T/V163A of the Green Fluorescent Protein (c3-GFP) has spectral characteristics similar to the wild-type GFP, but it is 42-fold more fluorescent in vivo. Here, we report the crystal structure and the refolding properties of c3-GFP and compare them with those of the less fluorescent wt-GFP and S65T mutant. The topology and the overall structure of c3-GFP is similar to the wild-type GFP. The three mutated residues, Ser99, Thr153, and Ala163, lie on the surface of the protein in three different beta-strands. The side chains of Ser99 and Thr153 are exposed to the solvent, whereas that of Ala163 points toward the interior of the protein. No significant deviation from the structure of the wild-type molecule is found around these positions, and there is not clear evidence of any distortion in the position of the chromophore or of the surrounding residues induced by the mutated amino acids. In vitro refolding experiments on urea-denatured c3-GFP reveal a renaturation behavior similar to that of the S65T molecule, with kinetic constants of the same order of magnitude. We conclude that the higher fluorescence activity of c3-GFP can be attributed neither to particular structural features nor to a faster folding process, as previously proposed.  相似文献   

2.
The site-directed mutations H82A and Q222A (residues near the heme access channel), and W171A and F267L (residues near the surface of the protein) were introduced into the gene encoding lignin peroxidase (LiP) isozyme H8 from Phanerochaete chrysosporium. The variant enzymes were produced by homologous expression in P. chrysosporium, purified to homogeneity, and characterized by kinetic and spectroscopic methods. The molecular masses, the pIs, and the UV-vis absorption spectra of the ferric and oxidized states of these LiP variant enzymes were similar to those of wild-type LiP (wtLiP), suggesting the overall protein and heme environments were not significantly affected by these mutations. The steady-state and transient-state parameters for the oxidation of veratryl alcohol (VA) by the H82A and Q222A variants were very similar to those of wtLiP, demonstrating that these residues are not involved in VA oxidation and that the heme access channel is an unlikely site for VA oxidation. In contrast, the W171A variant was unable to oxidize VA, confirming the apparent essentiality of Trp171 in VA oxidation by LiP. The kinetic rates of spontaneous LiP compound I reduction in the absence of VA were similar for W171A and wild-type LiP, suggesting that there may not be a radical formed on the Trp171 residue of LiP in the absence of VA. For the F267L variant, both the K(m app) value in the steady state and the apparent dissociation constant (K(D)) for compound II reduction were greater than those for wtLiP. These results indicate that the site including W171 and F267, rather than the heme access channel, is the site of VA binding and oxidation in LiP. Whereas Trp171 appears to be essential for VA oxidation, it apparently is not independently responsible for the spontaneous decomposition of oxidized intermediates. The nearby Phe267 apparently is also involved in VA binding.  相似文献   

3.
THE determination of sequences of histones F2A1(refs. 1 and 2) and F2B (ref. 3) and the amino-acid compositions of peptides of histone F1 (ref. 4) has shown that the distribution of residues along the polypeptide chains is extremely irregular. In particular, different regions of the molecule have markedly different characters. Thus the amino half of F2A1 has a ratio of basic to acidic residues of 15.6: 1 while for the carboxyl half of the molecule it is 1.5 : 1; for F2B this ratio is 13 : 1 for the amino quarter of the molecule and 1.9: 1 for the remainder. In the case of F1 this “polarity” is reversed: the carboxyl half of the molecule has a ratio of basic to acidic residues of 15 : 1, the amino half of 1.2 : 1. Further, regions rich in basic residues also contain a high proportion of helix destabilizing residues, proline in F1 and F2B and glycine in F2A1. In contrast, the non-basic regions contain a higher proportion of the apolar residues regarded as helix stabilizers and higher proportions of acidics, aromatics and threonine and serine residues. The characterization of two different regions led to the suggestion1–4 that the basic regions of the polypeptide chains are the primary sites of interaction with DNA, while the non-basic regions have the potential for the formation of definite conformations and might be involved in specific interactions other than with DNA.  相似文献   

4.
The replacement of Phe120 with other hydrophobic residues causes a decrease in the activity and thermal stability in ribonuclease A (RNase A). To explain this, the crystal structures of wild-type RNase A and three mutants--F120A, F120G, and F120W--were analyzed up to a 1.4 A resolution. Although the overall backbone structures of all mutant samples were nearly the same as that of wild-type RNase A, except for the C-terminal region of F120G with a high B-factor, two local conformational changes were observed at His119 in the mutants. First, His119 of the wild-type and F120W RNase A adopted an A position, whereas those of F120A and F120G adopted a B position, but the static crystallographic position did not reflect either the efficiency of transphosphorylation or the hydrolysis reaction. Second, His119 imidazole rings of all mutant enzymes were deviated from that of wild-type RNase A, and those of F120W and F120G appeared to be "inside out" compared with that of wild-type RNase A. Only approximately 1 A change in the distance between N(epsilon2) of His12 and N(delta1) of His119 causes a drastic decrease in k(cat), indicating that the active site requires the strict positioning of the catalytic residues. A good correlation between the change in total accessible surface area of the pockets on the surface of the mutant enzymes and enthalpy change in their thermal denaturation also indicates that the effects caused by the replacements are not localized but extend to remote regions of the protein molecule.  相似文献   

5.
The crystal structure of F65A/Y131C murine alpha-carbonic anhydrase V (CAV), covalently modified at cysteine residues with 4-chloromethylimidazole, is reported at 1.88 A resolution. This modification introduces a methylimidazole (MI) group at residue C131 in the active site with important consequences. F65A/Y131C-MI CAV exhibits an up to 3-fold enhancement of catalytic activity over that of wild-type CAV [Earnhardt, J. N., Wright, S. K., Qian, M., Tu, C., Laipis, P. J., Viola, R. E., and Silverman, D. N. (1999) Arch. Biochem. Biophys. 361, 264-270]. In this modified CAV variant, C131-MI acts as a proton shuttle, facilitating the deprotonation of a zinc-bound water molecule to regenerate the nucleophilic zinc-bound hydroxide ion. A network of three hydrogen-bonded water molecules, across which proton transfer likely proceeds, bridges the zinc-bound water molecule and the C131-MI imidazole group. The structure of F65A/Y131C-MI CAV is compared to structures of Y64H/F65A murine CAV, wild-type human alpha-carbonic anhydrase II, and the gamma-carbonic anhydrase from Methanosarcina thermophilain an effort to outline common features of catalytic proton shuttles.  相似文献   

6.
Previously, we have shown that residues 73-92 (sequence DRFSVNLDVKHFSPEELKVK) in alphaB-crystallin are involved in preventing the formation of light scattering aggregates by substrate proteins. In this study, we made single substitutions of three conserved amino acid residues (H83 --> A, F84 --> G, and P86 --> A) and a nonconserved amino acid residue (K90 --> C) in the functional region of alphaB-crystallin and evaluated their role in anti-aggregation activity. Mutation of conserved residues led to changes in intrinsic tryptophan intensity, bis-ANS binding, and in the secondary and tertiary structures. The H83A mutation led to a twofold increase in molar mass, while the other mutants did not produce significant changes in the molar mass when compared to that of wild-type protein. The chaperone-like activity of the H83A mutant was enhanced by 15%-20%, and the chaperone-like activity of F84G and P86A mutants was reduced by 50%-65% when compared to the chaperone-like activity of wild-type alphaB-crystallin. The substitution of the nonconserved residue (K90 --> C) did not induce an appreciable change in the structure and function of the mutant protein. Fluorescence resonance energy transfer (FRET) assay demonstrated that destabilized ADH interacted near the K90 region in alphaB-crystallin. The data show that F84 and P86 residues are essential for alphaB-crystallin to effectively prevent the aggregation of substrate proteins. This study further supports the involvement of the residues in the 73-92 region of alphaB-crystallin in substrate protein binding and chaperone-like action.  相似文献   

7.
8.
The Escherichia coli L-leucine receptor is an aqueous protein and the first component in the distinct transport pathway for hydrophobic amino acids. L-leucine binding induces a conformational change, which enables the receptor to dock to the membrane components. To investigate the ligand-induced conformational change and binding properties of this protein, we used (19)F NMR to probe the four tryptophan residues located in the two lobes of the protein. The four tryptophan residues were labeled with 5-fluorotryptophan and assigned by site-directed mutagenesis. The (19)F NMR spectra of the partially ligand free proteins show broadened peaks which sharpen when L-leucine is bound, showing that the labeled wild-type protein and mutants are functional. The titration of L-phenylalanine into the 5-fluorotryptophan labeled wild-type protein shows the presence of closed and open conformers. Urea-induced denaturation studies support the NMR results that the wild-type protein binds L-phenylalanine in a different manner to L-leucine. Our studies showed that the tryptophan to phenylalanine mutations on structural units linked to the binding pocket produce subtle changes in the environment of Trp18 located directly in the binding cleft.  相似文献   

9.
FliI is a Salmonella typhimurium protein that is needed for flagellar assembly and may be involved in a specialized protein export pathway that proceeds without signal peptide cleavage. FliI shows extensive sequence similarity to the catalytic beta subunit of the F0F1 ATPase (A. P. Volger, M. Homma, V. M. Irikura, and R. M. Macnab, J. Bacteriol. 173:3564-3572, 1991). It is even more similar to the Spa47 protein of Shigella flexneri (M. M. Venkatesan, J. M. Buysse, and E. V. Oaks, J. Bacteriol. 174:1990-2001, 1992) and the HrpB6 protein of Xanthomonas campestris (S. Fenselau, I. Balbo, and U. Bonas, Mol. Plant-Microbe Interact. 5:390-396, 1992), which are believed to play a role in the export of virulence proteins. Site-directed mutagenesis of residues in FliI that correspond to catalytically important residues in the F1 beta subunit resulted in loss of flagellation, supporting the hypothesis that FliI is an ATPase. FliI was overproduced and purified almost to homogeneity. It demonstrated ATP binding but not hydrolysis. An antibody raised against FliI permitted detection of the protein in wild-type cells and an estimate of about 1,500 subunits per cell. An antibody directed against the F1 beta subunit of Escherichia coli cross-reacted with FliI, confirming that the proteins are structurally related. The relationship between three proteins involved in flagellar assembly (FliI, FlhA, and FliP) and homologs in a variety of virulence systems is discussed.  相似文献   

10.
We have mapped the residues in the sixth transmembrane domains (TMs 6) of the mu, delta, and kappa opioid receptors that are accessible in the binding-site crevices by the substituted cysteine accessibility method (SCAM). We previously showed that ligand binding to the C7.38S mutants of the mu and kappa receptors and the wild-type delta receptor was relatively insensitive to methanethiosulfonate ethylammonium (MTSEA), a positively charged sulfhydryl-specific reagent. These MTSEA-insensitive constructs were used as the templates, and 22 consecutive residues in TM6 (excluding C6.47) of each receptor were mutated to cysteine, 1 at a time. Most mutants retained binding affinities for [3H]diprenorphine, a nonselective opioid antagonist, similar to that of the template receptors. Treatment with MTSEA significantly inhibited [3H]diprenorphine binding to 11 of 22 mutants of the rat mu receptor and 9 of 22 mutants of the human delta receptor and 10 of 22 mutants of the human kappa receptor. Naloxone or diprenorphine protected all sensitive mutants, except the A6.42(287)C mu mutant. Thus, V6.40, F6.44, W6.48, I6.51, Y6.54, V6.55, I6.56, I6.57, K6.58, and A6.59 of the mu receptor; F6.44, I6.51, F6.54, V6.55, I6.56, V6.57, W6.58, T6.59, and L6.60 of the delta receptor; and F6.44, W6.48, I6.51, F6.54, I6.55, L6.56, V6.57, E6.58, A6.59, and L6.60 of the kappa receptor are on the water-accessible surface of the binding-site crevices. The accessibility patterns of residues in the TMs 6 of the mu, delta, and kappa opioid receptors are consistent with the notion that the TMs 6 are in alpha-helical conformations with a narrow strip of accessibility on the intracellular side of 6.54 and a wider area of accessibility on the extracellular side of 6.54, likely due to a proline kink at 6.50 that bends the helix in toward the binding pocket and enables considerable motion in this region. The wider exposure of residues 6.55-6.60 to the binding-site crevice, combined with the divergent amino acid sequences, is consistent with the inferred role of residues in this region in determining ligand binding selectivity. The conservation of the accessibility pattern on the cytoplasmic side of 6.54 suggests that this region may be important for receptor activation. This accessibility pattern is similar to that of the D2 dopamine receptor, the only other GPCR in which TM6 has been mapped by SCAM. That opioid receptors and the remotely related D2 dopamine receptor have similar accessibility patterns in TM6 suggest that these segments of GPCRs in the rhodopsin-like subfamily not only share secondary structure but also are packed similarly into the transmembrane bundle and thus have similar tertiary structure.  相似文献   

11.
Metallothioneins have been isolated from Saccharomyces cerevisiae CUP1 mutants generated by Wright et al. (Wright, C. F., Hamer, D. H., and McKenney, K. (1986) Nucleic Acids Res. 14, 8489-8499). In the mutant metallothioneins, pairs of cysteinyl residues have been converted to seryl residues. The mutant proteins differ only in the positions of the double substitutions; each mutant molecule contains 10 cysteinyl residues. Each mutant protein lacks the first 8 residues at the amino terminus from the decoded gene sequence of the CUP1 locus. Mutant molecules consist of 53 residues analogous to the wild-type metallothionein and are designated 9/11, 24/26, 36/38, and 49/50 (in reference to the sequence positions of the Cys----Ser conversions). The properties of the mutant metallothioneins are vastly different, and host cells harboring the different plasmid-encoded mutant molecules show marked differences in sensitivity to CuSO4. Growth inhibition was observed at CuSO4 concentrations up to mM in cells containing the 9/11, 24/26, and 36/38 molecules, but not for cells containing protein 49/50. A CuSO4 concentration of 5 mM was required to inhibit the growth of yeast containing either 49/50 or the wild-type metallothionein. In the purified proteins the copper binding stoichiometry of each molecule, except protein 24/26, was nearly 8 mol eq. Protein 24/26 bound 5.5 copper ions/molecule. The Cu(I) chelator bathocuproine disulfonate reacted with over 50% of the copper ions in proteins 9/11, 24/26, and 36/38, but less than 10% of the copper ions in proteins 49/50 and wild-type metallothionein were reactive. The thiolates in 9/11, 24/26, and 36/38 were also more reactive in a disulfide exchange reaction with dithiodipyridine compared with the sulfhydryls in 49/50 and the wild-type molecules. The four mutant copper proteins are luminescent and exhibit a similar quantum yield. The cluster structures contributing to the particular electronic transitions are markedly more sensitive to oxygen in proteins 9/11, 24/26, and 36/38 compared with 49/50 and the wild-type molecules. The air-sensitive proteins exhibit a tertiary fold not recognized by polyclonal antibodies directed to a conformational epitope on yeast Cu-metallothionein. Protein 49/50 cross-reacts with the antibody in a concentration-dependent fashion similar to the wild-type protein. Mutation of 2 cysteinyl residues in the carboxyl portion of metallothionein does not significantly alter properties of the molecule, whereas mutation of several cysteines in the amino-terminal portion of the molecule yields a different conformation.  相似文献   

12.
Phe286 located in the center of the active site of alpha-amylase 2 from Thermoactinomyces vulgaris R-47 (TVAII) plays an important role in the substrate recognition for cyclomaltooligosaccharides (cyclodextrins). The X-ray structures of mutant TVAIIs with the replacement of Phe286 by Ala (F286A) and Tyr (F286Y) were determined at 3.2 A resolution. Their structures have no significant differences from that of the wild-type enzyme. The kinetic analyses of Phe286-replaced variants showed that the variants with non-aromatic residues, Ala (F286A) and Leu (F286L), have lower enzymatic activities than those with aromatic residues, Tyr (F286Y) and Trp (F286W), and the replacement of Phe286 affects enzymatic activities for CDs more than those for starch.  相似文献   

13.
THE amino-acid sequences of the histones F2A11 and F2B2 are of particular interest in that they have such a high degree of non-uniformity that different regions of the polypeptide chains have quite different characters. The amino-halves of the molecules have a high density of basic residues while the carboxyl-halves have far fewer basic residues and higher proportions of apolar residues and other residues which favour the formation of helical conformations. These properties have led to the suggestions that the regions of high basicity are the primary sites for interaction with DNA1–4 in chromatin and that the non-basic regions contain any secondary structure that the molecule is capable of forming1–4 and are the sites for histone–histone interactions3,4. Evidence in support of this scheme has been obtained from nuclear magnetic resonance and optical spectroscopic studies of conformational changes and interactions in histones F1 and F2A13 and F2B4. In these studies, however, the whole molecule has been examined and the possibility of the formation of some secondary structure in the basic region of the molecule cannot be excluded.  相似文献   

14.
The Phe46 residue, located in the hydrophobic core of RNase A, was replaced with other hydrophobic residues, leucine, valine, or alanine, and their X-ray crystallographic structures were determined up to 1.50-1.80 A resolution in an attempt to examine the relationship between structural changes and conformational stability or folding kinetics. The backbone structure of F46L, F46V, and F46A was indistinguishable from that of the wild-type enzyme, retaining the correct active site structure. However, one water molecule was included in the hydrophobic core of F46A, forming two hydrogen bonds with the backbone peptide chain. The side chain of Met29 in F46V and F46A adopted two different conformations in an equal occupancy. A trapped water molecule and two conformations of Met29 represent changes that minimize the cavity volume. Nevertheless, the replacement of Phe46 with the above residues resulted in a marked decrease in both thermal stability and folding reaction. Thus, Phe46 ensures the thermal stability and the rapid and correct folding of RNase A by the role it plays in forming a highly packed, hydrophobic core.  相似文献   

15.
SER virus, a paramyxovirus that is closely related to simian virus 5 (SV5), is unusual in that it fails to induce syncytium formation. The SER virus F protein has an unusually long cytoplasmic tail (CT), and it was previously observed that truncations or specific mutations of this domain result in enhanced syncytium formation. In addition to the long CT, the SER F protein has nine amino acid differences from the F protein of SV5. We previously observed only a partial suppression of fusion in a chimeric SV5 F protein with a CT derived from SER virus, indicating that these other amino acid differences between the SER and SV5 F proteins also play a role in regulating the fusion phenotype. To examine the effects of individual amino acid differences, we mutated the nine SER residues individually to the respective residues of the SV5 F protein. We found that most of the mutants were expressed well and were transported to the cell surface at levels comparable to that of the wild-type SER F protein. Many of the mutants showed enhanced lipid mixing, calcein transfer, and syncytium formation even in the presence of the long SER F protein CT. Some mutants, such as the I310 M, T438S, M489I, T516V, and N529K mutants, also showed fusion at lower temperatures of 32, 25, and 18 degrees C. The residue Asn529 plays a critical role in the suppression of fusion activity, as the mutation of this residue to lysine caused a marked enhancement of fusion. The effect of the N529K mutation on the enhancement of fusion by a previously described mutant, L539,548A, as well as by chimeric SV5/SER F proteins was also dramatic. These results indicate that activation to a fusogenic conformation is dependent on the interplay of residues in the ectodomain, the transmembrane domain, and the CT domain of paramyxovirus F proteins.  相似文献   

16.
19F nuclear magnetic resonance (19F NMR) of 5-fluorotryptophan (5F-Trp) and tryptophan (Trp) fluorescence both provide information about local environment and solvent exposure of Trp residues. To compare the information provided by these spectroscopies, the four Trp residues in recombinant soluble human tissue factor (sTF) were replaced with 5F-Trp. 19F NMR assignments for the 5F-Trp residues (14, 25, 45, and 158) were based on comparison of the wild-type protein spectrum with the spectra of three single Trp-to-Phe replacement mutants. Previously we showed from fluorescence and absorption difference spectra of mutant versus wild-type sTF that the side chains of Trpl4 and Trp25 are buried, whereas those of Trp45 and Trp158 are partially exposed to bulk solvent (Hasselbacher et al., Biophys J 1995;69:20-29). 19F NMR paramagnetic broadening and solvent-induced isotope-shift experiments show that position 5 of the indole ring of 5F-Trp158 is exposed, whereas that of 5F-Trp45 is essentially inaccessible. Although 5F-Trp incorporation had no discernable effect on the procoagulant cofactor activity of either the wild-type or mutant proteins, 19F NMR chemical shifts showed that the single-Trp mutations are accompanied by subtle changes in the local environments of 5F-Trp residues residing in the same structural domain.  相似文献   

17.
We cloned a bovine cDNA encoding the neural adhesion molecule F3 and analyzed its nucleotide sequence. The coding region consisted of 3054 bp encoding 1018 amino acid (aa) residues. The Mr calculated from the deduced as sequence was 113 383. Bovine F3 had 93, 94 and 77% as identity with the mouse, human and chicken homologs, respectively. Bovine F3, similar to those of chicken and human, was devoid of two as residues (Ile-Thr) in the sixth immunoglobulin type C2-like domain, as compared with the mouse homolog. Parts of bovine F3 protein were overproduced in Escherichia coli. The antibodies raised against the recombinant proteins in rabbits reacted specifically with F3. F3 protein was detected in cerebellum, cerebrum and spinal cord in Western blot analysis.  相似文献   

18.
The thermo-stability and unfolding behaviors of a small hyperthermophilic protein Sso7d as well as its single-point mutation F31A are studied by molecular dynamics simulation at temperatures of 300 K, 371 K and 500 K. Simulations at 300 K show that the F31A mutant displays a much larger flexibility than the wild type, which implies that the mutation obviously decreases the protein's stability. In the simulations at 371 K, although larger fluctuations were observed, both of these two maintain their stable conformations. High temperature simulations at 500 K suggest that the unfolding of these two proteins evolves along different pathways. For the wild-type protein, the C-terminal alpha-helix is melted at the early unfolding stage, whereas it is destroyed much later in the unfolding process of the F31A mutant. The results also show that the mutant unfolds much faster than its parent protein. The deeply buried aromatic cluster in the F31A mutant dissociates quickly relative to the wild-type protein at high temperature. Besides, it is found that the triple-stranded antiparallel β-sheet in the wild-type protein plays an important role in maintaining the stability of the entire structure.  相似文献   

19.
20.
A thermophilic Bacillus stearothermophilus F1 produces an extremely thermostable serine protease. The F1 protease sequence was used to predict its three-dimensional (3D) structure to provide better insights into the relationship between the protein structure and biological function and to identify opportunities for protein engineering. The final model was evaluated to ensure its accuracy using three independent methods: Procheck, Verify3D, and Errat. The predicted 3D structure of F1 protease was compared with the crystal structure of serine proteases from mesophilic bacteria and archaea, and led to the identification of features that were related to protein stabilization. Higher thermostability correlated with an increased number of residues that were involved in ion pairs or networks of ion pairs. Therefore, the mutants W200R and D58S were designed using site-directed mutagenesis to investigate F1 protease stability. The effects of addition and disruption of ion pair networks on the activity and various stabilities of mutant F1 proteases were compared with those of the wild-type F1 protease.  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