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.
The crystal structure of the gene V protein (GVP) from the Ff filamentous phages (M13, fl, fd) has been solved for the wild-type and two mutant (Y41F and Y41H) proteins at high resolution. The Y41H mutant crystal structure revealed crystal packing interactions, which suggested a plausible scheme for constructing the polymeric protein shell of the GVP-single-stranded DNA (ssDNA) complex (Guan Y, et al., 1994, Biochemistry 33:7768-7778). The electrostatic potentials of the isolated and the cooperatively formed protein shell have been calculated using the program GRASP and they revealed a highly asymmetric pattern of the electrostatic charge distribution. The inner surface of the putative DNA-binding channel is positively charged, whereas the opposite outer surface is nearly neutral. The electrostatic calculation further demonstrated that the formation of the helical protein shell enhanced the asymmetry of the electrostatic distribution. A model of the GVP-ssDNA complex with the n = 4 DNA-binding mode could be built with only minor conformational perturbation to the GVP protein shell. The model is consistent with existing biochemical and biophysical data and provides clues to the properties of GVP, including the high cooperatively of the protein binding to ssDNA. The two antiparallel ssDNA strands form a helical ribbon with the sugar-phosphate backbones at the middle and the bases pointing away from each other. The bases are stacked and the Phe 73 residue is intercalated between two bases. The optimum binding to a tetranucleotide unit requires the participation of four GVP dimers, which may explain the cooperativity of the GVP binding to DNA.  相似文献   

2.
3.
The gene 5 protein (g5p) of the Ff virus contains five Tyr, individual mutants of which have now all been characterized by CD spectroscopy. The protein has a dominant tyrosyl 229-nm L(a) CD band that is shown to be approximately the sum of the five individual Tyr contributions. Tyr41 is particularly important in contributing to the high cooperativity with which the g5p binds to ssDNA, and Y41F and Y41H mutants are known to differ in dimer-dimer packing interactions in crystal structures. We compared the solution structures and binding properties of the Y41F and Y41H mutants using CD spectroscopy. Secondary structures of the mutants were similar by CD analyses and close to those derived from the crystal structures. However, there were significant differences in the binding properties of the two mutant proteins. The Y41H protein had an especially low binding affinity and perturbed the spectrum of poly[d(A)] in 2 mM Na(+) much less than did Y41F and the wild-type gene 5 proteins. Moreover, a change in the Tyr 229 nm band, assigned to the perturbation of Tyr34 at the dimer-dimer interface, was absent in titrations with the Y41H mutant under low salt conditions. In contrast, titrations with the Y41H mutant in 50 mM Na(+) exhibited typical CD changes of both the nucleic acid and the Tyr 229-nm band. Thus, protein-protein and g5p-ssDNA interactions appeared to be mutually influenced by ionic strength, indicative of correlated changes in the ssDNA binding and cooperativity loops of the protein or of indirect structural constraints.  相似文献   

4.
Gas vesicles of cyanobacteria are formed by a protein called 'gas-vesicle protein' (GVP). The complete amino acid sequence has been determined of GVP from Anabaena flos-aquae. It is 70 residues long and has an Mr of 7388. This corresponds to the size of the repeating unit cell demonstrated by X-ray crystallography of intact gas vesicles. Details of the sequence are related to the secondary beta-sheet structure of the protein and its contrasting hydrophilic and hydrophobic surfaces. Extensive amino acid sequences have also been determined for GVPs from two other cyanobacteria, species of Calothrix and Microcystis; they are highly homologous with that of Anabaena GVP. Electrophoretic analysis indicates that GVPs of different cyanobacteria form a variety of stable oligomers.  相似文献   

5.
Mutational analysis of the disulfide catalysts DsbA and DsbB   总被引:2,自引:0,他引:2  
In prokaryotes, disulfides are generated by the DsbA-DsbB system. DsbB functions to generate disulfides by quinone reduction. These disulfides are passed to the DsbA protein and then to folding proteins. To investigate the DsbA-DsbB catalytic system, we performed an in vivo selection for chromosomal dsbA and dsbB mutants. We rediscovered many residues previously shown to be important for the activity of these proteins. In addition, we obtained one novel DsbA mutant (M153R) and four novel DsbB mutants (L43P, H91Y, R133C, and L146R). We also mutated residues that are highly conserved within the DsbB family in an effort to identify residues important for DsbB function. We found classes of mutants that specifically affect the apparent K(m) of DsbB for either DsbA or quinones, suggesting that quinone and DsbA may interact with different regions of the DsbB protein. Our results are consistent with the interpretation that the residues Q33 and Y46 of DsbB interact with DsbA, Q95 and R48 interact with quinones, and that residue M153 of DsbA interacts with DsbB. All of these interactions could be due to direct amino acid interactions or could be indirect through, for instance, their effect on protein structure. In addition, we find that the DsbB H91Y mutant severely affects the k(cat) of the reaction between DsbA and DsbB and that the DsbB L43P mutant is inactive, suggesting that both L43 and H91 are important for the activity of DsbB. These experiments help to better define the residues important for the function of these two protein-folding catalysts.  相似文献   

6.
7.
Kobayashi T  Ikeguchi M  Sugai S 《Proteins》2002,49(3):297-301
At neutral pH, equine beta-lactoglobulin (ELG) is monomeric, whereas bovine beta-lactoglobulin (BLG) exists as a dimer. To understand the difference in the oligomerization properties between ELG and BLG, three mutants of ELG (LP, I, and LPI) were constructed by substituting amino acids responsible for important interactions at the dimer interface of BLG into ELG. The mutant LP has an AB loop mutation (S34A/E35Q), the mutant I has an I strand mutation (G145M/R146H/V147I/Q148R/I149L/V150S/P151F/D152N/L153P) and the mutant LPI includes both the LP and I mutations. The far- and near-UV CD spectra of the three mutants are similar to that of the wild-type ELG, indicating that the secondary and the tertiary structures of ELG are not significantly affected by the mutations. Ultracentrifuge analysis shows that all three mutants are monomeric at neutral pH, suggesting that the protein sequences in the AB loop and I strand of BLG alone cannot support dimerization of ELG. Thus, structural differences must exist between ELG and BLG that prevent the ELG mutants from forming the same interactions as BLG at the dimer interface.  相似文献   

8.
We have recently shown an essential role of the 32 amino acids C-terminus domain of IcmT of Legionella pneumophila in bacterial egress from macrophages. Mutants expressing an IcmT protein with a truncation in the C-terminus, replicate intracellularly but are defective in pore formation-mediated egress. The C-terminus domain of IcmT is the only hydrophilic domain of IcmT that is predicted to be in the cytoplasm while the rest of the protein is in the cytoplasmic membrane. In order to characterize the structure-function of the C-terminus of IcmT in the pore-forming activity and bacterial egress, we constructed 10 icmT missense mutant alleles differing by a single amino acid in the C-terminus of icmT and introduced them into the null icmT mutant. The H58Q, W69L, R71I, R79I and R86I icmT mutant alleles showed significantly lower pore-forming activity as measured by hemolysis of sRBC. The Y59S, R68L and S77L mutant alleles showed significantly lower cytopathogenicity to U937 macrophages. All 10 mutant alleles enabled the icmT null mutant to replicate intracellularly as efficiently as icmT null mutant harboring the wild-type icmT. Seven of the icmT alleles enabled the icmT null mutant to egress from infected macrophages as efficiently as icmT null mutant harboring the wild-type icmT. The other 3 substitutions conferred a partial defect in hemolysis and two of them also conferred a defect in egress from macrophages. Thus, two amino acid residues in the C-terminus of IcmT are required for both pore formation and bacterial egress. However, certain single amino acid substitutions in the C-terminus reduce the pore-forming activity when tested in vitro, but may or may not have a detectable effect on egress of L. pneumophila from U937 macrophages.  相似文献   

9.
Reaction centers from the Y(L167) mutant of Rhodobacter sphaeroides, containing a highly oxidizing bacteriochlorophyll dimer and a tyrosine residue substituted at Phe L167, were compared to reaction centers from the Y(M) mutant, with a tyrosine at M164, and a quadruple mutant containing a highly oxidizing dimer but no nearby tyrosine residue. Distinctive features in the light-induced optical and EPR spectra showed that the oxidized bacteriochlorophyll dimer was reduced by Tyr L167 in the Y(L167) mutant, resulting in a tyrosyl radical, as has been found for Tyr M164 in the Y(M) mutant. In the Y(L167) mutant, the net proton uptake after formation of the tyrosyl radical and the reduced primary quinone ranged from +0.1 to +0.3 H(+)/reaction center between pH 6 and pH 10, with a dependence that is similar to the quadruple mutant but different than the large proton release observed in the Y(M) mutant. In the light-induced absorption spectrum in the 700-1000 nm region, the Y(L167) mutant exhibited unique changes that can be assigned as arising primarily from an approximately 30 nm blue shift of the dimer absorption band. The optical signals in the Y(L167) mutant were pH dependent, with a pK(a) value of approximately 8.7, indicating that the tyrosyl radical is stabilized at high pH. The results are modeled by assuming that the phenolic proton of Tyr L167 is trapped in the protein after oxidation of the tyrosine, resulting in electrostatic interactions with the tetrapyrroles and nearby residues.  相似文献   

10.
11.
This investigation describes how the binding characteristics of the single-stranded DNA-binding protein encoded by gene V of bacteriophage M13, are affected by single-site amino acid substitutions. The series of mutant proteins tested includes mutations in the purported monomer-monomer interaction region as well as mutations in the DNA-binding domain at positions which are thought to be functionally involved in monomer-monomer interaction or single-stranded DNA binding. The characteristics of the binding of the mutant proteins to the homopolynucleotides poly(dA), poly(dU) and poly(dT), were studied by means of fluorescence-titration experiments. The binding stoichiometry and fluorescence quenching of the mutant proteins are equal to, or lower than, the wild-type gene V protein values. In addition, all proteins measured bind a more-or-less co-operative manner to single-stranded DNA. The binding affinities for poly(dA) decrease in the following order: Y61H greater than wild-type greater than F68L and R16H greater than Y41F and Y41H greater than F73L greater than R21C greater than Y34H greater than G18D/Y56H. Possible explanations for the observed differences are discussed. The conservation of binding affinity, also for mutations in the single-stranded DNA-binding domain, suggests that the binding to homopolynucleotides is largely non-specific.  相似文献   

12.
XRCC1 operates as a scaffold protein in base excision repair, a pathway that copes with base and sugar damage in DNA. Studies using recombinant XRCC1 proteins revealed that: a C389Y substitution, responsible for the repair defects of the EM-C11 CHO cell line, caused protein instability; a V86R mutation abolished the interaction with POLβ, but did not disrupt the interactions with PARP-1, LIG3α and PCNA; and an E98K substitution, identified in EM-C12, reduced protein integrity, marginally destabilized the POLβ interaction, and slightly enhanced DNA binding. Two rare (P161L and Y576S) and two frequent (R194W and R399Q) amino acid population variants had little or no effect on XRCC1 protein stability or the interactions with POLβ, PARP-1, LIG3α, PCNA or DNA. One common population variant (R280H) had no pronounced effect on the interactions with POLβ, PARP-1, LIG3α and PCNA, but did reduce DNA-binding ability. When expressed in HeLa cells, the XRCC1 variants—excluding E98K, which was largely nucleolar, and C389Y, which exhibited reduced expression—exhibited normal nuclear distribution. Most of the protein variants, including the V86R POLβ-interaction mutant, displayed normal relocalization kinetics to/from sites of laser-induced DNA damage: except for E98K and C389Y, and the polymorphic variant R280H, which exhibited a slightly shorter retention time at DNA breaks.  相似文献   

13.
14.
The complete amino acid sequence of coagulogen purified from the hemocytes of the horseshoe crab Carcinoscorpius rotundicauda was determined by characterization of the NH2-terminal sequence and the peptides generated after digestion of the protein with lysyl endopeptidase, Staphylococcal aureus protease V8 and trypsin. Upon sequencing the peptides by the automated Edman method, the following sequence was obtained: A D T N A P L C L C D E P G I L G R N Q L V T P E V K E K I E K A V E A V A E E S G V S G R G F S L F S H H P V F R E C G K Y E C R T V R P E H T R C Y N F P P F V H F T S E C P V S T R D C E P V F G Y T V A G E F R V I V Q A P R A G F R Q C V W Q H K C R Y G S N N C G F S G R C T Q Q R S V V R L V T Y N L E K D G F L C E S F R T C C G C P C R N Y Carcinoscorpius coagulogen consists of a single polypeptide chain with a total of 175 amino acid residues and a calculated molecular weight of 19,675. The secondary structure calculated by the method of Chou and Fasman reveals the presence of an alpha-helix region in the peptide C segment (residue Nos. 19 to 46), which is released during the proteolytic conversion of coagulogen to coagulin gel. The beta-sheet structure and the 16 half-cystines found in the molecule appear to yield a compact protein stable to acid and heat. The amino acid sequences of coagulogen of four species of limulus have been compared and the interspecies evolutionary differences are discussed.  相似文献   

15.
Enzymes that salvage 6-oxopurines, including hypoxanthine phosphoribosyltransferases (HPRTs), are potential targets for drugs in the treatment of diseases caused by protozoan parasites. For this reason, a number of high-resolution X-ray crystal structures of the HPRTs from protozoa have been reported. Although these structures did not reveal why HPRTs need to form dimers for catalysis, they revealed the existence of potentially relevant interactions involving residues in a loop of amino acid residues adjacent to the dimer interface, but the contributions of these interactions to catalysis remained poorly understood. The loop, referred to as active-site loop I, contains an unusual non-proline cis-peptide and is composed of residues that are structurally analogous with Leu67, Lys68, and Gly69 in the human HPRT. Functional analyses of site-directed mutations (K68D, K68E, K68N, K68P, and K68R) in the HPRT from Trypanosoma cruzi, etiologic agent of Chagas' disease, show that the side-chain at position 68 can differentially influence the K(m) values for all four substrates as well as the k(cat) values for both IMP formation and pyrophosphorolysis. Also, the results for the K68P mutant are inconsistent with a cis-trans peptide isomerization-assisted catalytic mechanism. These data, together with the results of structural studies of the K68R mutant, reveal that the side-chain of residue 68 does not participate directly in reaction chemistry, but it strongly influences the relative efficiencies for IMP formation and pyrophosphorolysis, and the prevalence of lysine at position 68 in the HPRT of the majority of eukaryotes is consistent with there being a biological role for nucleotide pyrophosphorolysis.  相似文献   

16.
Green fluorescent protein (GFP) fluoresces efficiently under blue excitation despite major electrostatic rearrangements resulting from photoionization of the chromophore and neutralization of Glu-222. A competing phototransformation process, which ionizes the chromophore and decarboxylates Glu-222, mimics the electrostatic and structural changes in the fluorescence photocycle. Structural and spectroscopic analysis of the cryogenically stabilized photoproduct at 100 K and a structurally annealed intermediate of the phototransformed protein at 170 K reveals distinct structural relaxations involving protein, chromophore, solvent, and photogenerated CO2. Strong structural changes of the 100 K photoproduct after decarboxylation appear exclusively within 15 angstroms of the chromophore and include the electrostatically driven perturbations of Gln-69, Cys-70, and water molecules in an H-bonding network connecting the chromophore. X-ray crystallography to 1.85 angstroms resolution and static and picosecond time-resolved IR spectroscopy identify structural mechanisms common to phototransformation and to the fluorescence photocycle. In particular, the appearance of a 1697 cm(-1) (+) difference band in both photocycle and phototransformation intermediates is a spectroscopic signature for the structural perturbation of Gln-69. This is taken as evidence for an electrostatically driven dynamic response that is common to both photoreaction pathways. The interactions between the chromophore and the perturbed residues and solvent are decreased or removed in the T203H single and T203H/Q69L double mutants, resulting in a strong reduction of the fluorescence quantum yield. This suggests that the electrostatic response to the transient formation of a buried charge in the wild type is important for the bright fluorescence.  相似文献   

17.
Feng J  Goswami S  Howell EE 《Biochemistry》2008,47(2):555-565
R67 dihydrofolate reductase (DHFR) bears no sequence or structural homologies with chromosomal DHFRs. The gene for this enzyme produces subunits that are 78 amino acids long, which assemble into a homotetramer possessing 222 symmetry. More recently, a tandem array of four gene copies linked in-frame was constructed, which produces a monomer containing 312 amino acids named Quad3. Asymmetric mutations in Quad3 have also been constructed to probe the role of Q67 and K32 residues in catalysis. This present study mixes and matches mutations to determine if the Q67H mutation, which tightens binding approximately 100-fold to both dihydrofolate (DHF) and NADPH, can help rescue the K32M mutation. While the latter mutation weakens DHF binding over 60-fold, it concurrently increases kcat by a factor of 5. Two Q67H mutations were added to gene copies 1 and 4 in conjunction with the K32M mutation in gene copies 1 and 3. Addition of these Q67H mutations tightens binding 40-fold, and the catalytic efficiency (kcat/Km(DHF)) of the resulting protein is similar to that of Quad3. Since these Q67H mutations can mostly compensate for the K32M lesion, K32 must not be necessary for DHF binding. Another multimutant combines the K32M mutation in gene copies 1 and 3 with the Q67H mutation in all gene copies. This mutant is inhibited by DHF but not NADPH, indicating that NADPH binds only to the wild type half of the pore, while DHF can bind to either the wild type or mutant half of the pore. This inhibition pattern contrasts with the mutant containing only the Q67H substitution in all four gene copies, which is severely inhibited by both NADPH and substrate. Since gene duplication and divergence are evolutionary tools for gaining function, these constructs are a first step toward building preferences for NADPH and DHF in each half of the active site pore of this primitive enzyme.  相似文献   

18.
The 0.5beta monoclonal antibody is a very potent strain-specific HIV-neutralizing antibody raised against gp120, the envelope glycoprotein of HIV-1. This antibody recognizes the V3 loop of gp120, which is a major neutralizing determinant of the virus. The antibody-peptide interactions, involving aromatic and negatively charged residues of the antibody 0.5beta, were studied by NMR and double-mutant cycles. A deuterated V3 peptide and a Fab containing deuterated aromatic amino acids were used to assign these interactions to specific V3 residues and to the amino acid type and specific chain of the antibody by NOE difference spectroscopy. Electrostatic interactions between negatively charged residues of the antibody Fv and peptide residues were studied by mutagenesis of both antibody and peptide residues and double-mutant cycles. Several interactions could be assigned unambiguously: F96(L) of the antibody interacts with Pro13 of the peptide, H52(H) interacts with Ile7, Ile9 and Gln10 and D56(H) interacts with Arg11. The interactions of the light-chain tyrosines with Pro13 and Gly14 could be assigned to either Y30a(L) and Y32(L), respectively, or Y32(L) and Y49(L), respectively. Three heavy-chain tyrosines interact with Ile7, Ile20 and Phe17. Several combinations of assignments involving Y32(H), Y53(H), Y96(H) and Y100a(H) may satisfy the NMR and mutagenesis constraints, and therefore at this stage the interactions of the heavy-chain tyrosines were not taken into account. The unambiguous assignments [F96(L), H52(H) and D56(H)] and the two possible assignments of the light-chain tyrosines were used to dock the peptide into the antibody-combining site. The peptide converges to a unique position within the binding site, with the RGPG loop pointing into the center of the groove formed by the antibody complementary determining regions while retaining the beta-hairpin conformation and the type-VI RGPG turn [Tugarinov, V., Zvi, A., Levy, R. & Anglister, J. (1999) Nat. Struct. Biol. 6, 331-335].  相似文献   

19.
Bartish G  Nygård O 《Biochimie》2008,90(5):736-748
Elongation factor 2 (eEF2) is a member of the G-protein super family. G-proteins undergo conformational changes associated with binding of the guanosine nucleotide and hydrolysis of the bound GTP. These structural rearrangements affects the Switch I region (also known as the Effector loop). We have studied the role of individual amino acids in the Switch I region (amino acids 25-73) of S. cerevisiae eEF2 using functional complementation in yeast. 21 point mutations in the Switch I region were created by site-directed mutagenesis. Mutants K49R, E52Q, A53G, F55Y, K60R, Q63A, T68S, I69M and A73G were functional while mutants R54H, F55N, D57A, D57E, D57S, R59K, R59M, Q63E, R65A, R65N, T68A and T68M were inactive. Expression of mutants K49R, A53G, Q63A, I69M and A73G was associated with markedly decreased growth rates and yeast cells expressing mutants A53G and I69M became temperature sensitive. The functional capacity of eEF2 in which the major part Switch I (amino acids T56 to I69) was converted into the homologous sequence found in EF-G from E. coli was also studied. This protein chimera could functionally replace yeast eEF2 in vivo. Yeast cells expressing this mutant grew extremely slowly, showed increased cell death and became temperature sensitive. The ability of the mutant to replace authentic eEF2 in vivo indicates that the structural rearrangement of Switch I necessary for eEF2 function is similar in eukaryotes and bacteria. The effect of two point mutations in the P-loop was also studied. Mutant A25G but not A25V could functionally replace yeast eEF2 even if cells expressing the mutant grew slowly. The A25G mutation converted the consensus sequences AXXXXGK[T/S] in eEF2 to the corresponding motif GXXXXGK[T/S] found in all other G-proteins, suggesting that the alanine found in the P-loop of peptidyltranslocases are not essential for function.  相似文献   

20.
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