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.
β-Lactoglobulin (LG) contains nine β-strands (strands A-I) and one α-helix. Strands A-H form a β-barrel. At neutral pH, bovine LG (BLG) forms a dimer and the dimer interface consists of AB-loops and the I-strands of two subunits. On the other hand, equine LG (ELG) is monomeric. The residues 145-153 of BLG, which compose a dimer interface, are entirely different from those of ELG. The difference in the association states between BLG and ELG can be attributed to the residues 145-153. To confirm this, we constructed a chimeric LG, ImBLG (I-strand mutated BLG), in which the residues 145-153 were replaced with those of ELG. Gel-filtration chromatography and analytical ultracentrifugation revealed that ImBLG existed as a monomer. To identify the residues important for dimerization, we constructed several revertants and investigated their association. This experiment revealed that, in addition to the interface residues (Ile147, Leu149 and Phe151), Met145 is critical for dimerization. Although Met145 does not contact with the other protomer, it seems to be important in determining the backbone conformation of the I-strand. This was supported by the fact that all Met145-containing mutants showed circular dichroism spectra similar to BLG but different from ImBLG.  相似文献   

2.
β-lactoglobulin (LG) contains nine β-strands (strands A-I) and one α-helix. Strands A-H form a β-barrel. At neutral pH, equine LG (ELG) is monomeric, whereas bovine LG (BLG) is dimeric, and the I-strands of its two subunits form an intermolecular β-sheet. We previously constructed a chimeric ELG in which the sequence of the I-strand was replaced with that of BLG. This chimera did not dimerize. For this study, we constructed the new chimera we call Gyuba (which means cow and horse in Japanese). The amino acid sequence of Gyuba includes the sequences of the BLG secondary structures and those of the ELG loops. The crystal structure of Gyuba is very similar to that of BLG and indicates that Gyuba dimerizes via the intermolecular β-sheet formed by the two I-strands. Thus, the entire arrangement of the secondary structural elements is important for LG dimer formation.  相似文献   

3.
Acidic buffer conditions are known to stabilize helix-rich states of even those proteins with a predominantly beta-sheet native secondary structure. Here we investigated whether such states also exist under alkaline buffer conditions. The guanidine hydrochloride (GuHCl)-induced unfolding transition and kinetic refolding of equine beta-lactoglobulin (ELG) by GuHCl-jump were investigated at pH 8.7 by far-ultraviolet circular dichroism. We found that an equilibrium intermediate appeared in 45% ethylene glycol (EGOH) buffer with 1.5 M GuHCl. The intermediate is rich in non-native alpha-helix, which is similar to the helix-rich state of ELG at pH 4.0. A kinetic study was done on the folding rate of ELG and compared with bovine beta-lactoglobulin (BLG). Transient intermediates, which were observed as the burst phase of the refolding reaction, were also rich in alpha-helix. The activation enthalpy of ELG was calculated to be c.a. 80 kJ/mol, whereas that of BLG was c.a. 70 kJ/mol in the presence of 45% EGOH. The ellipticities of the transient intermediate of ELG show temperature dependence in the presence of 45% EGOH, whereas that of BLG did not show significant dependence. This study therefore extends the existence of helix-rich equilibrium and transient intermediates of predominantly beta-sheet proteins to alkaline buffer conditions.  相似文献   

4.
The anthranilate phosphoribosyltransferase from Sulfolobus solfataricus (ssAnPRT) forms a homodimer with a hydrophobic subunit interface. To elucidate the role of oligomerisation for catalytic activity and thermal stability of the enzyme, we loosened the dimer by replacing two apolar interface residues with negatively charged residues (mutations I36E and M47D). The purified double mutant I36E+M47D formed a monomer with wild-type catalytic activity but reduced thermal stability. The single mutants I36E and M47D were present in a monomer-dimer equilibrium with dissociation constants of about 1 μM and 20 μM, respectively, which were calculated from the concentration-dependence of their heat inactivation kinetics. The monomeric form of M47D, which is populated at low subunit concentrations, was as thermolabile as monomeric I36E+M47D. Likewise, the dimeric form of I36E, which was populated at high subunit concentrations, was as thermostable as dimeric wild-type ssAnPRT. These findings show that the increased stability of wild-type ssAnPRT compared to the I36E+M47D double mutant is not caused by the amino acid exchanges per se but by the higher intrinsic stability of the dimer compared to the monomer. In accordance with the negligible effect of the mutations on catalytic activity and stability, the X-ray structure of M47D contains only minor local perturbations at the dimer interface. We conclude that the monomeric double mutant resembles the individual wild-type subunits, and that ssAnPRT is a dimer for stability but not for activity reasons.  相似文献   

5.
Mammalian glutathione (GSH) transferases are dimeric proteins, many of which share a common hydrophobic interaction motif that is important for dimer stability. In the rGSTM1-1 enzyme this motif involves the side chain of F56, located on the 56 loop of the N-terminal domain, which is intercalated between the alpha4- and alpha5-helices of the C-terminal domain of the opposing subnuit. Disruption of the complementary interactions in this motif by mutation of F56 to serine, arginine, or glutamate is known to have deleterious effects on catalytic efficiency but remarkably different effects on the stability of the dimer [Hornby et al. (2002) Biochemistry 41, 14238-14247]. The structural basis for the behavior of the mutants by amide H/D exchange mass spectrometry is described. A substantial decrease in H/D exchange is observed in the GSH binding domain and in parts of the dimer interface upon substrate binding. The F56S and F56R mutants exhibit enhanced H/D exchange kinetics in the GSH binding domain and at the dimer interface. In contrast, the F56E mutant shows a decrease in the rate and extent of amide H/D exchange at the dimer interface and enhanced exchange kinetics in the GSH binding domain. The results suggest that the F56E mutant has a restructured dimer interface with decreased solvent accessibility and dynamics. Although all of the F56 mutations disrupt the GSH binding site, the effects of the mutations on the structure of the subunit interface and dimer stability are quite distinct.  相似文献   

6.
The lipocalin beta-lactoglobulin (BLG) is the major whey protein of bovine milk and is homodimeric at physiological conditions. Each monomer contains two disulfide bonds and one cysteine at position 121 (C121). This free thiol plays an important role in the heat-induced aggregation of BLG and, possibly, in its conformational stability. We describe here the expression in the yeast Pichia pastoris of a mutant bovine BLG, in which C121 was changed into Ser (C121S). Circular dichroism and high-performance liquid chromatography experiments, together with the X-ray crystal structure, show that the C121S mutant retains a nativelike fold at both neutral and acid pH. The mutation completely blocks the irreversible aggregation induced by heat treatment at 90 degrees C. Compared to the recombinant wild-type protein, the mutant is less stable to temperature and disulfide reducing agents and is much more sensitive to peptic digestion. Moreover, its affinity for 1-anilino-8-naphthalenesulfonate is increased at neutral and acid pH. We suggest that the stability of the protein arising from the hydrophobic effect is reduced by the C121S mutation so that unfolded or partially unfolded states are more favored.  相似文献   

7.
Bovine beta-lactoglobulin (BLG) is a globular protein of uncertain physiological function and a member of the lipocalin superfamily of proteins. Here, we present the X-ray structure at 3.0 angstroms of BLG (variant A) from an orthorhombic (P2(1)2(1)2(1)) pseudo-tetragonal crystal form that suffers from pseudo-merohedral twinning (final R(working) = 0.224, R(free) = 0.265). Crystals were grown by dialysis against ultra-purified water (i.e., at very low ionic strength), at pH approximately 5.2 (approximately pI), conditions vastly different from all other BLG structures determined previously. This allows critical assessment of the BLG structure and of the influence that pH, ionic strength, and crystal packing may have on the molecular structure of BLG. The pH-sensitive EF loop is found in the closed conformation characteristic of BLG at pH less than 7 and moderate to high ionic strength. Although the hydrophobic pocket appears to be empty, the presence of highly disordered water molecules cannot be excluded. The dimer interface and the hydrophobic pocket (calyx) are preserved. However, the orientation of the subunits in the dimer varies considerably with crystal form. Structure is deposited with PDB ID 2akq.  相似文献   

8.
The monomer-dimer equilibrium for the human immunodeficiency virus type 1 (HIV-1) protease has been investigated under physiological conditions. Dimer dissociation at pH 7.0 was correlated with a loss in beta-sheet structure and a lower degree of ANS binding. An autolysis-resistant mutant, Q7K/L33I/L63I, was used to facilitate sedimentation equilibrium studies at neutral pH where the wild-type enzyme is typically unstable in the absence of bound inhibitor. The dimer dissociation constant (KD) of the triple mutant was 5.8 microM at pH 7.0 and was below the limit of measurement (approximately 100 nM) at pH 4.5. Similar studies using the catalytically inactive D25N mutant yielded a KD value of 1.0 microM at pH 7.0. These values differ significantly from a previously reported value of 23 nM obtained indirectly from inhibitor binding measurements (Darke et al., 1994). We show that the discrepancy may result from the thermodynamic linkage between the monomer-dimer and inhibitor binding equilibria. Under conditions where a significant degree of monomer is present, both substrates and competitive inhibitors will shift the equilibrium toward the dimer, resulting in apparent increases in dimer stability and decreases in ligand binding affinity. Sedimentation equilibrium studies were also carried out on several drug-resistant HIV-1 protease mutants: V82F, V82F/I84V, V82T/I84V, and L90M. All four mutants exhibited reduced dimer stability relative to the autolysis-resistant mutant at pH 7.0. Our results indicate that reductions in drug affinity may be due to the combined effects of mutations on both dimer stability and inhibitor binding.  相似文献   

9.
Electrospray ionization mass spectrometry (ESI MS) under nanospray conditions has been used to examine the effects of mutation at two key dimer interface residues, Gln (Q) 64 and Thr (T) 75, in Plasmodium falciparum triosephosphate isomerase. Both residues participate in an intricate network of intra- and intersubunit hydrogen bonds. The gas phase distributions of dimeric and monomeric protein species have been examined for the wild type enzyme (TWT) and three mutants, Q64N, Q64E, and T75S, under a wide range of collision energies (40–160 eV). The results established the order of dimer stability as TWT > T75S > Q64E ∼ Q64N. The mutational effects on dimer stability are in good agreement with the previously reported estimates, based on the concentration dependence of enzyme activity. Additional experiments in solution, using inhibition of activity by a synthetic dimer interface peptide, further support the broad agreement between gas phase and solution studies.  相似文献   

10.
Catalysis of DNA recombination by Tn3 resolvase is conditional on prior formation of a synapse, comprising 12 resolvase subunits and two recombination sites (res). Each res binds a resolvase dimer at site I, where strand exchange takes place, and additional dimers at two adjacent 'accessory' binding sites II and III. 'Hyperactive' resolvase mutants, that catalyse strand exchange at site I without accessory sites, were selected in E. coli. Some single mutants can resolve a res x site I plasmid (that is, with one res and one site I), but two or more activating mutations are necessary for efficient resolution of a site I x site I plasmid. Site I x site I resolution by hyperactive mutants can be further stimulated by mutations at the crystallographic 2-3' interface that abolish activity of wild-type resolvase. Activating mutations may allow regulatory mechanisms of the wild-type system to be bypassed, by stabilizing or destabilizing interfaces within and between subunits in the synapse. The positions and characteristics of the mutations support a mechanism for strand exchange by serine recombinases in which the DNA is on the outside of a recombinase tetramer, and the tertiary/quaternary structure of the tetramer is reconfigured.  相似文献   

11.
Green AL  Hrodey HA  Brooker RJ 《Biochemistry》2003,42(38):11226-11233
Previous work on the lactose permease of Escherichia coli has shown that mutations along a face of predicted transmembrane segment 8 (TMS-8) play a critical role in conformational changes associated with lactose transport (Green, A. L., and Brooker, R. J. [2001] Biochemistry 40, 12220-12229). Substitutions at positions 261, 265, 268, 272, and 276, which form a continuous stripe along TMS-8, were markedly defective for lactose transport velocity. In the current study, three single mutants (F261D, N272Y, N272L) and a double mutant (T265Y/M276Y) were chosen as parental strains for the isolation of mutants that restored transport function. A total of 68 independent mutants were isolated and sequenced. Forty-four were first-site revertants in which the original mutation was changed back to the wild-type residue or to a residue with a similar side-chain volume. The other 24 mutations were second-site suppressors in TMS-2 (Q60L, Q60P), loop 2/3 (L70H), TMS-7 (V229G/A), TMS-8 (F261L), and TMS-11 (F354V, C355G). On the basis of their locations, the majority of the second-site suppressors can be interpreted as improving the putative TMS-2/TMS-7/TMS-11 interface to compensate for conformational defects imposed by mutations in TMS-8 that disrupt the putative TMS-1/TMS-5/TMS-8 interface. Overall, this paper suggests that the TMS-2/TMS-7/TMS-11 interface is more important from a functional point of view, even though there is compelling evidence for structural symmetry between the two halves of the permease.  相似文献   

12.
The aggregation of α-synuclein is linked directly to the histopathology of Parkinson’s disease (PD). However, several missense mutations present in the α-synuclein gene (SNCA) have been known to be associated with PD. Several studies have highlighted the effect of SNCA mutations on the α-synuclein aggregation, but their pathological roles are not completely established. In this study, we have focused on the effects of the recently discovered α-synuclein missense mutants (H50Q and G51D) on the aggregation using computational approaches. We performed all atom molecular dynamics (MD) simulation on these mutants and compared their conformational dynamics with Wild-Type (WT) α-synuclein. We noticed the solvent accessible surface area (SASA), radius of gyration, atomic fluctuations, and beta strand content to be higher in H50Q than G51D and WT. Using PDBSum online server; we analyzed the inter-molecular interactions that drive the association of monomeric units of H50Q, WT, and G51D in forming the respective homo-dimer. We noticed the interface area, number of interacting residues and binding free energy to be higher for H50Q homo-dimer than the WT and G51D homo-dimers. Our findings in this study suggest that in comparison to WT and G51D, H50Q mutation to have a positive effect on increasing the α-synuclein aggregation propensity. Hence, we see that H50Q and G51D mutation show conflicting effect on the aggregation propensity of α-synuclein.  相似文献   

13.
Point mutations were introduced into the major capsid protein (P3) of cloned infectious cDNA of the polerovirus beet western yellows virus (BWYV) by manipulation of cloned infectious cDNA. Seven mutations targeted sites on the S domain predicted to lie on the capsid surface. An eighth mutation eliminated two arginine residues in the R domain, which is thought to extend into the capsid interior. The effects of the mutations on virus capsid formation, virus accumulation in protoplasts and plants, and aphid transmission were tested. All of the mutants replicated in protoplasts. The S-domain mutant W166R failed to protect viral RNA from RNase attack, suggesting that this particular mutation interfered with stable capsid formation. The R-domain mutant R7A/R8A protected approximately 90% of the viral RNA strand from RNase, suggesting that lower positive-charge density in the mutant capsid interior interfered with stable packaging of the complete strand into virions. Neither of these mutants systemically infected plants. The six remaining mutants properly packaged viral RNA and could invade Nicotiana clevelandii systemically following agroinfection. Mutant Q121E/N122D was poorly transmitted by aphids, implicating one or both targeted residues in virus-vector interactions. Successful transmission of mutant D172N was accompanied either by reversion to the wild type or by appearance of a second-site mutation, N137D. This finding indicates that D172 is also important for transmission but that the D172N transmission defect can be compensated for by a "reverse" substitution at another site. The results have been used to evaluate possible structural models for the BWYV capsid.  相似文献   

14.
Point mutations at the dimer interface of the homodimeric enzyme ascorbate peroxidase (APx) were constructed to assess the role of quaternary interactions in the stability and activity of APx. Analysis of the APx crystal structure shows that Glu112 forms a salt bridge with Lys20 and Arg24 of the opposing subunit near the axis of dyad symmetry between the subunits. Two point mutants, E112A and E112K, were made to determine the effects of a neutral (alanine) and repulsive (lysine) mutation on dimerization, stability, and activity. Gel filtration analysis indicated that the ratio of the monomer to dimer increased as the dimer interface interactions went from attractive to repulsive. Differential scanning calorimetry (DSC) data exhibited a decrease in both the transition temperature (Tm) and enthalpy of unfolding (deltaHc) with Tm = 58.3 +/- 0.5 degrees C, 56.0 +/- 0.8 degrees C, and 53.0 +/- 0.9 degrees C and deltaHc = 245 +/- 29 kcal/mol, 199 +/- 38 kcal/mol, and 170 +/- 25 kcal/mol for wild-type APx, E112A, and E112K, respectively. Similar changes were observed based on thermal melting curves obtained by absorption spectroscopy. No change in enzyme activity was found for the E112A mutant, and only a 25% drop in activity was observed for the E112K mutant which demonstrates that the non-Michaelis Menten kinetics of APx is not due to the APx oligomeric structure. The cryogenic crystal structures of the wild-type and mutant proteins show that mutation induced changes are limited to the dimer interface including an alteration in solvent structure.  相似文献   

15.
Human cystatin C (HCC) is one of the amyloidogenic proteins to be shown to oligomerize via a three‐dimensional domain swapping mechanism. This process precedes the formation of a stable dimer and proceeds particularly easily in the case of the L68Q mutant. According to the proposed mechanism, dimerization of the HCC precedes conformational changes within the β2 and β3 strands. In this article, we present conformational studies, using circular dichroism and MD methods, of the β2‐L1‐β3 (His43‐Thr72) fragment of the HCC involved in HCC dimer formation. We also carried out studies of the β2‐L1‐β3 peptide, in which the Val57 residue was replaced by residues promoting β‐turn structure formation (Asp, Asn, or Pro). The present study established that point mutation could modify the structure of the L1 loop in the β‐hairpin peptide. Our results showed that the L1 loop in the peptide excised from human cystatin C is broader than that in cystatin C. In the HCC protein, broadening of the L1 loop together with the unfavorable L68Q mutation in the hydrophobic pocket could be a force sufficient to cause the partial unfolding and then the opening of HCC or its L68Q mutant structure for further dimerization. We presume further that the Asp57 and Asn57 mutations in the L1 loop of HCC could stabilize the closed form of HCC, whereas the Pro57 mutation could lead to the opening of the HCC structure and then to dimer/oligomer formation. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 373–383, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

16.
Fabry disease is an X-linked lysosomal storage disorder caused by the deficiency of alpha-galactosidase A that results in the accumulation of neutral sphingolipids. We report a novel point mutation in exon 6, Q279K, carried by an asymptomatic child with a family history of classic Fabry disease. Moreover, we comparatively study the in vitro expression and enzyme activity of Q279K and three other already described mutants in glutamine 279. The Q279K, Q279H and Q279R mutants transfected in COS-1 cells expressed no activity while the residual enzyme activity of the Q279E mutant represented 10% of wild type value. Western blot analysis demonstrated a differential behavior of the mutant proteins: Q279K and Q279H persisted as the inactive 50-kD precursor, indicating that these mutations may affect the normal processing of the enzyme, while the Q279R mutant was not detected probably due to an unstable protein which is rapidly degraded. The in vitro expression studies of the novel Q279K mutation were confirmed by Western blot analysis performed in the patient's lymphocytes which revealed the alpha-galactosidase A precursor of 50 kD but not the processed form.  相似文献   

17.
Canonical glutathione (GSH) transferases are dimeric proteins with subunits composed of an N-terminal GSH binding region (domain 1) and a C-terminal helical region (domain 2). The stabilities of several GSH transferase dimers are dependent upon two groups of interactions between domains 1 and 2 of opposing subunits: a hydrophobic ball-and-socket motif and a buried charge cluster motif. In rGSTM1-1, these motifs involve residues F56 and R81, respectively. The structural basis for the effects of mutating F56 to different residues on dimer stability and function has been reported (Codreanu et al. (2005) Biochemistry 44, 10605-10612). Here, we show that the simultaneous disruption of both motifs in the F56S/R81A mutant causes complete dissociation of the dimer to a monomeric protein on the basis of gel filtration chromatography and multiple-angle laser light scattering. The fluorescence and far-UV CD properties of the double mutant as well as the kinetics of amide H/D exchange along the polypeptide backbone suggest that the monomer has a globular structure that is similar to a single subunit in the native protein. However, the mutant monomer has severely impaired catalytic activity, suggesting that the dimer interface is vital for efficient catalysis. Backbone amide H/D exchange kinetics in the F56S and F56S/R81A mutants indicate that a reorganization of the loop structure between helix alpha2 and strand beta3 near the active site is responsible for the decreased catalytic activity of the monomer. In addition, the junction between the alpha4 and alpha5 helices in F56S/R81R shows decreased H/D exchange, indicating another structural change that may affect catalysis. Although the native subunit interface is important for dimer stability, urea-induced unfolding of the F56S/R81A mutant suggests that the interface is not essential for the thermodynamic stability of individual subunits. The H/D exchange data reveal a possible molecular basis for the folding cooperativity observed between domains 1 and 2.  相似文献   

18.
19.
ClC-1 is a dimeric, double-pored chloride channel that is present in skeletal muscle. Mutations of this channel can result in the condition myotonia, a muscle disorder involving increased muscle stiffness. It has been shown that the dominant form of myotonia often results from mutations that affect the so-called slow, or common, gating process of the ClC-1 channel. Mutations causing dominant myotonia are seen to cluster at the interface of the ClC-1 channel monomers. This study has investigated the role of the H, I, P, and Q helices, which lie on this interface, as well as the G helix, which is situated immediately behind the H and I helices, on ClC-1 gating. 11 mutant ClC-1 channels (T268M, C277S, C278S, S289A, T310M, S312A, V321S, T539A, S541A, M559T, and S572V) were produced using site-directed mutagenesis, and gating properties of these channels were investigated using electrophysiological techniques. Six of the seven mutations in G, H, and I, and two of the four mutations in P and Q, caused shifts of the ClC-1 open probability. In the majority of cases this was due to alterations in the common gating process, with only three of the mutants displaying any change in fast gating. Many of the mutant channels also showed alterations in the kinetics of the common gating process, particularly at positive potentials. The changes observed in common gating were caused by changes in the opening rate (e.g. T310M), the closing rate (e.g. C277S), or both rates. These results indicate that mutations in the helices forming the dimer interface are able to alter the ClC-1 common gating process by changing the energy of the open and/or closed channel states, and hence altering transition rates between these states.  相似文献   

20.
To explore the influence of the I(L177)H single mutation on the properties of the nearest bacteriochlorophylls (BChls), three reaction centers (RCs) bearing double mutations were constructed in the photosynthetic purple bacterium Rhodobacter sphaeroides, and their properties and pigment content were compared with those of the correspondent single mutant RCs. Each pair of the mutations comprised the amino acid substitution I(L177)H and another mutation altering histidine ligand of BChl P(A) or BChl B(B). Contrary to expectations, the double mutation I(L177)H+H(L173)L does not bring about a heterodimer RC but causes a 46nm blue shift of the long-wavelength P absorbance band. The histidine L177 or a water molecule were suggested as putative ligands for P(A) in the RC I(L177)H+H(L173)L although this would imply a reorientation of the His backbone and additional rearrangements in the primary donor environment or even a repositioning of the BChl dimer. The crystal structure of the mutant I(L177)H reaction center determined to a resolution of 2.9? shows changes at the interface region between the BChl P(A) and the monomeric BChl B(B). Spectral and pigment analysis provided evidence for β-coordination of the BChl B(B) in the double mutant RC I(L177)H+H(M182)L and for its hexacoordination in the mutant reaction center I(L177)H. Computer modeling suggests involvement of two water molecules in the β-coordination of the BChl B(B). Possible structural consequences of the L177 mutation affecting the coordination of the two BChls P(A) and B(B) are discussed. This article is part of a Special Issue entitled: Photosynthesis Research for Sustainability: from Natural to Artificial.  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