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.
The amino acid Pro is more rigid than other naturally occurring amino acids and, in proteins, lacks an amide hydrogen. To understand the structural and thermodynamic effects of Pro substitutions, it was introduced at 13 different positions in four different proteins, leucine-isoleucine-valine binding protein, maltose binding protein, ribose binding protein, and thioredoxin. Three of the maltose binding protein mutants were characterized by X-ray crystallography to confirm that no structural changes had occurred upon mutation. In the remaining cases, fluorescence and CD spectroscopy were used to show the absence of structural change. Stabilities of wild type and mutant proteins were characterized by chemical denaturation at neutral pH and by differential scanning calorimetry as a function of pH. The mutants did not show enhanced stability with respect to chemical denaturation at room temperature. However, 6 of the 13 single mutants showed a small but significant increase in the free energy of thermal unfolding in the range of 0.3-2.4 kcal/mol, 2 mutants showed no change, and 5 were destabilized. In five of the six cases, the stabilization was because of reduced entropy of unfolding. However, the magnitude of the reduction in entropy of unfolding was typically several fold larger than the theoretical estimate of -4 cal K(-1) mol(-1) derived from the relative areas in the Ramachandran map accessible to Pro and Ala residues, respectively. Two double mutants were constructed. In both cases, the effects of the single mutations on the free energy of thermal unfolding were nonadditive.  相似文献   

2.
Properties of mutationally altered RNA polymerases II of Drosophila   总被引:9,自引:0,他引:9  
We tested and compared several in vitro properties of wild type and mutant RNA polymerases II from Drosophila melanogaster, using several different mutants of a single X-linked genetic locus, RpIIC4 (Greenleaf, A. L., Weeks, J. R., Voelker, R. A., Ohnishi, S., and Dickson, B. (1980) Cell 21, 785-792); the mutants tested included the original amanitin-resistant mutant, C4, which is nonconditional, plus the temperature-sensitive mutants A9, C20, E28, and 1Fb40. Using a tritium-labeled amanitin derivative, we demonstrated that C4 polymerase has a reduced binding affinity for amanitin. The C4 polymerase was as stable to thermal denaturation as the wild type enzyme, and the two enzymes had similar specific activities, ionic strength and Mn2+ requirements, and apparent Km values for UTP and GTP when assayed in the presence of Mn2+. However, with Mg2+ as the divalent cation, C4 polymerase was less active than wild type and had 2-fold higher apparent Km values for UTP and GTP. Three of the temperature-sensitive mutants, A9, C20, and E28, were derived from the amanitin-resistant mutant C4; the polymerase II activities from these mutants displayed resistance to alpha-amanitin in vitro identical with that of the C4 enzyme. C20, E28, and 1Fb40 polymerases were markedly less stable to thermal denaturation in vitro than wild type polymerase. The results presented indicate that the mutations at the RNA polymerase locus (RpIIC4-) directly alter the structure of the enzyme, providing conclusive evidence that the locus is a structural gene for a polymerase II subunit.  相似文献   

3.
MICA is a major histocompatibility complex-like protein that undergoes a structural transition from disorder to order upon binding its immunoreceptor, NKG2D. We redesigned the disordered region of MICA with RosettaDesign to increase NKG2D binding. Mutations that stabilize this region were expected to increase association kinetics without changing dissociation kinetics, increase affinity of interaction, and reduce entropy loss upon binding. MICA mutants were stable in solution, and they were amenable to surface plasmon resonance evaluation of NKG2D binding kinetics and thermodynamics. Several MICA mutants bound NKG2D with enhanced affinity, kinetic changes were primarily observed during association, and thermodynamic changes in entropy were as expected. However, none of the 15 combinations of mutations predicted to stabilize the receptor-bound MICA conformation enhanced NKG2D affinity, whereas all 10 mutants predicted to be destabilized bound NKG2D with increased on-rates. Five of these had affinities enhanced by 0.9-1.8 kcal/mol over wild type by one to three non-contacting substitutions. Therefore, in this case, mutations designed to mildly destabilize a protein enhanced association and affinity.  相似文献   

4.
Cross FR  Levine K 《Genetics》2000,154(4):1549-1559
We showed recently that a screen for mutant CDC28 with improved binding to a defective Cln2p G1 cyclin yielded a spectrum of mutations similar to those yielded by a screen for intragenic suppressors of the requirement for activation loop phosphorylation (T169E suppressors). Recombination among these mutations yielded CDC28 mutants that bypassed the G1 cyclin requirement. Here we analyze further the interrelationship between T169E suppression, interaction with defective cyclin, and G1 cyclin bypass. DNA shuffling of mutations from the various screens and recombination onto a T169E-encoding 3' end yielded CDC28 mutants with strong T169E suppression. Some of the strongest T169E suppressors could suppress the defective Cln2p G1 cyclin even while retaining T169E. The strong T169E suppressors did not exhibit bypass of the G1 cyclin requirement but did so when T169E was reverted to T. These results suggested that for these mutants, activation loop phosphorylation and cyclin binding might be alternative means of activation rather than independent requirements for activation (as with wild type). These results suggest mechanistic overlap between the conformational shift induced by cyclin binding and that induced by activation loop phosphorylation. This conclusion was supported by analysis of suppressors of a mutation in the Cdk phosphothreonine-binding pocket created by cyclin binding.  相似文献   

5.
Mutational study of the bacterial hemoglobin distal heme pocket   总被引:1,自引:0,他引:1  
Ligand binding experiments on three mutants in the distal heme pocket of Vitreoscilla hemoglobin (GlnE7His, ProE8Ala, and GlnE7His,ProE8Ala) were used to probe the role of GlnE7 and ProE8 in the pocket's unusual structure. The oxygen dissociation constants for the wild type, E8Ala mutant, and E7His mutant proteins were 4.5, 4.7, and 1.7microM, respectively; the K(d) for the double mutant was not determinable by our technique. Visible-Soret spectra of the carbonyl and cyanyl forms and FT-IR of the carbonyl form of the E8 mutant were similar to those of the wild type; the opposite was true for the GlnE7His and GlnE7His,ProE8Ala mutants, which also differed from wild type in the visible-Soret spectra of their oxidized forms. Models of the effects of the mutations on distal pocket structure were consistent with the experimental findings, particularly the larger effects of the GlnE7His change.  相似文献   

6.
The structural properties of myristoylated forms of recombinant recoverin of the wild type and of its mutants with damaged second and/or third Ca(2+)-binding sites were studied by fluorimetry and circular dichroism. The interaction of wild-type recoverin with calcium ions was shown to induce unusual structural rearrangements in its molecule. In particular, protein binding with Ca2+ ions results in an increase in the mobility of the environment of Trp residues, in higher hydrophobicity, and in elevated thermal stability (its thermal transition shifts by 15 degrees C to higher temperatures) but has almost no effect on its secondary structure. Similar structural changes induced by Ca2+ are also characteristic of the -EF2 mutant of recoverin whose second Ca(2+)-binding site is modified and cannot bind calcium ions. The structural properties of the -EF3 and -EF2,3 mutants (whose third or simultaneously second and third Ca(2+)-binding sites, respectively, are modified and damaged) are practically indifferent to calcium ions.  相似文献   

7.
We have analyzed four Nicotiana plumbaginifolia null mutants presumably affected in the heme domain of nitrate reductase. The DNA sequence of this domain has been determined for each mutant and for the wild type. Two mutations were identified as single base changes leading to, respectively, the substitution of a histidine residue by an asparagine (mutant E56) and to the appearance of an ochre stop codon (mutant E64). Based on the amino acid sequence homology between the nitrate reductase heme domain and mammalian cytochrome b5, we have predicted the three-dimensional structure of this domain. This showed that the nitrate reductase heme domain is structurally very similar to cytochrome b5 and it also confirmed that the residue involved in E56 mutation is one of the two heme-binding histidines. The two other mutations (mutants A1 and K21) were found to be, respectively, -1 and +1 frameshift mutations resulting in the appearance of an opal stop codon. These sequence data confirmed previous genetic and biochemical hypotheses on nitrate reductase-deficient mutants. Northern blot analysis of these mutants indicated that mutant E56 overexpressed the nitrate reductase mRNA, whereas the nonsense mutations present in the other mutants led to reduced levels of nitrate reductase mRNA.  相似文献   

8.
The glyceraldehyde-3-phosphate dehydrogenase (GAPDH) in the chloroplast of Chlamydomonas reinhardtii is part of a complex that also includes phosphoribulokinase (PRK) and CP12. We identified two residues of GAPDH involved in protein-protein interactions in this complex, by changing residues K128 and R197 into A or E. K128A/E mutants had a Km for NADH that was twice that of the wild type and a lower catalytic constant, whatever the cofactor. The kinetics of the mutant R197A were similar to those of the wild type, while the R197E mutant had a lower catalytic constant with NADPH. Only small structural changes near the mutation may have caused these differences, since circular dichroism and fluorescence spectra were similar to those of wild-type GAPDH. Molecular modelling of the mutants led to the same conclusion. All mutants, except R197E, reconstituted the GAPDH-CP12 subcomplex. Although the dissociation constants measured by surface plasmon resonance were 10-70-fold higher with the mutants than with wild-type GAPDH and CP12, they remained low. For the R197E mutation, we calculated a 4 kcal/mol destabilizing effect, which may correspond to the loss of the stabilizing effect of a salt bridge for the interaction between GAPDH and CP12. All the mutant GAPDH-CP12 subcomplexes failed to interact with PRK and to form the native complex. The absence of kinetic changes of all the mutant GAPDH-CP12 subcomplexes, compared to wild-type GAPDH-CP12, suggests that mutants do not undergo the conformation change essential for PRK binding.  相似文献   

9.
Protein misfolding due to missense mutations is a common pathogenic mechanism in cystathionine β-synthase (CBS) deficiency. In our previous studies, we successfully expressed, purified, and characterized nine CBS mutant enzymes containing the following patient mutations: P49L, P78R, A114V, R125Q, E176K, R266K, P422L, I435T, and S466L. These purified mutants exhibited full heme saturation, normal tetrameric assembly, and high catalytic activity. In this work, we used several spectroscopic and proteolytic techniques to provide a more thorough insight into the conformation of these mutant enzymes. Far-UV circular dichroism, fluorescence, and second-derivative UV spectroscopy revealed that the spatial arrangement of these CBS mutants is similar to that of the wild type, although the microenvironment of the chromophores may be slightly altered. Using proteolysis with thermolysin under native conditions, we found that the majority of the studied mutants is more susceptible to cleavage, suggesting their increased local flexibility or propensity for local unfolding. Interestingly, the presence of the CBS allosteric activator, S-adenosylmethionine (AdoMet), increased the rate of cleavage of the wild type and the AdoMet-responsive mutants, while the proteolytic rate of the AdoMet-unresponsive mutants was not significantly changed. Pulse proteolysis analysis suggested that the protein structure of the R125Q and E176K mutants is significantly less stable than that of the wild type and the other mutants. Taken together, the proteolytic data shows that the conformation of the pathogenic mutants is altered despite retained catalytic activity and normal tetrameric assembly. This study demonstrates that the proteolytic techniques are useful tools for the assessment of the biochemical penalty of missense mutations in CBS.  相似文献   

10.
We applied in vitro mutagenesis and colony screening, using the wild type phyI1s gene from Aspergillus niger 113 as the template, and obtained two mutant phyI1s (gene products) after one round of screening. The two mutants had mutations at two nucleic acid sites, resulting in changes in two amino acids: K41E, E121F. None of the amino acid substitutions in the two mutants was in a position reported to be important for catalysis or substrate binding. Kinetic analysis of the phytase activity of the two mutants indicated that the substitutions gave rise to 2.5- and 3.1-fold increased specific activity, and a 1.78- and 3.24-fold reduced affinity for sodium phytate. In addition, the overall catalytic efficiency (k cat/K m) of the two mutants was changed by 0.52-fold and 0.68-fold compared to that of the wild type. Such mutants will be instrumental for the structure–function study of the enzyme and for industrial application.  相似文献   

11.
The effects of three mutations on the EF-hand Ca(2+)/Mg(2+) binding site of smooth muscle myosin regulatory light chain (RLC) were studied: D5S, in which an aspartate is replaced by a serine in position 5 of the loop; D9E, in which an aspartate is replaced by a glutamate in position 9; and D12E, in which the aspartate in position 12 is replaced by a glutamate. All possible combinations of the three mutations were produced. The single mutants D5S and D9E and the double mutant D5S/D9E have low affinity for Ca(2+). All the mutants containing mutation D12E are Ca(2+)-specific and have higher affinities than wild type, even when containing mutations D5S or D9E. All of the mutants studied have lower affinity for Mg(2+) than the wild-type protein. As expected, the changes in binding free energy that each mutant produces depend on the residues present at the other positions of the site, since the mutated positions are very close in the protein structure. Coupling energies are about the same for all pairs of mutants when binding Ca(2+), but can have different values when binding Mg(2+). D5S and D9E have a large negative coupling energy for Mg(2+) binding which suggests an interaction between these two positions. When mutation D12E is present, the coupling energy for Mg(2+) binding between D5S and D9E is much lower, suggesting that this interaction occurs only if an aspartate is in position 12. Glutamate in position 9 may be able to coordinate Mg(2+) directly in the double mutant D5S/D9E.  相似文献   

12.
Protein-protein association is accompanied by a large reduction in translational and rotational (external) entropy. Based on a 15 ns molecular dynamics simulation of acetylcholinesterase (AChE) in complex with fasciculin 2 (Fas2), we estimate the loss in external entropy using quasiharmonic analysis and histogram-based approximations of the probability distribution function. The external entropy loss of AChE-Fas2 binding, ~30 cal/mol K, is found to be significantly larger than most previously characterized protein-ligand systems. However, it is less than the entropy loss estimated in an earlier study by A. V. Finkelstein and J. Janin, which was based on atomic motions in crystals.  相似文献   

13.
Calmodulin (CaM) modulates the activity of several proteins that play a key role in excitation-contraction coupling (ECC). In cardiac muscle, the major binding partner of CaM is the type-2 ryanodine receptor (RyR2) and altered CaM binding contributes to defects in sarcoplasmic reticulum (SR) calcium (Ca2+) release. Many genetic studies have reported a series of CaM missense mutations in patients with a history of severe arrhythmogenic cardiac disorders. In the present study, we generated four missense CaM mutants (CaMN98I, CaMD132E, CaMD134H and CaMQ136P) and we used a CaM-RyR2 co-immunoprecipitation and a [3H]ryanodine binding assay to directly compare the relative RyR2-binding of wild type and mutant CaM proteins and to investigate the functional effects of these CaM mutations on RyR2 activity. Furthermore, isothermal titration calorimetry (ITC) experiments were performed to investigate and compare the interactions of the wild-type and mutant CaM proteins with various synthetic peptides located in the well-established RyR2 CaM-binding region (3584-3602aa), as well as another CaM-binding region (4255-4271aa) of human RyR2. Our data revealed that all four CaM mutants displayed dramatically reduced RyR2 interaction and defective modulation of [3H]ryanodine binding to RyR2, regardless of LQTS or CPVT association. Moreover, our isothermal titration calorimetry ITC data suggest that RyR2 3584-3602aa and 4255-4271aa regions interact with significant affinity with wild-type CaM, in the presence and absence of Ca2+, two regions that might contribute to a putative intra-subunit CaM-binding pocket. In contrast, screening the interaction of the four arrhythmogenic CaM mutants with two synthetic peptides that correspond to these RyR2 regions, revealed disparate binding properties and signifying differential mechanisms that contribute to reduced RyR2 association.  相似文献   

14.
Death of Saccharomyces cerevisiae induced by sulfur dioxide (K(2)S(2)O(2) was used as the SO(2) source) followed saturation kinetics. The enthalpy of activation of death was not affected by concentration over the range tested (5-150) mg/L of (K(2)S(2)O(2) at pH 3.4) and averaged 3.6 x 10(4) cal/mol as compared with 8.5 x 10(4) cal/mol for DeltaH(double dagger) of thermal death. The entropy of activation of death was hyperbolic function of the sulfur dioxide concentration, extrapolated at zero concentration to DeltaS(0) (double dagger) = 36.8 cal mol(-1) K(-1) and tended to DeltaDeltaS(max) (double dagger) = 13.2 cal mol(-1) K(-1) at saturating concentration, yielding a dis0sociation constant of 5.8 x 10(-1) M sulfur dioxide. As was predicted from these results, In K(d) (the specific rate of death induced by sulfur dioxide) was hyperbolic function of concentration under isothermic conditions and extrapolated toa finite value at zero concentration. The Arrhenuis plots and the DeltaS(+/-) plot versus concentration revealed the occurrence of substrate inhabitation of the death effect at high concentrations (above 60 mg/L K(2)S(2)O(2) at pH 3.4). A model is presented involving two types of receptor sites for sulfur dioxide on the cell surface, on directly connected with the death process, the other modulating its entropy of activation.  相似文献   

15.
Harris TK  Wu G  Massiah MA  Mildvan AS 《Biochemistry》2000,39(7):1655-1674
The MutT enzyme catalyzes the hydrolysis of nucleoside triphosphates (NTP) to NMP and PP(i) by nucleophilic substitution at the rarely attacked beta-phosphorus. The solution structure of the quaternary E-M(2+)-AMPCPP-M(2+) complex indicated that conserved residues Glu-53, -56, -57, and -98 are at the active site near the bound divalent cation possibly serving as metal ligands, Lys-39 is positioned to promote departure of the NMP leaving group, and Glu-44 precedes helix I (residues 47-59) possibly stabilizing this helix which contributes four catalytic residues to the active site [Lin, J. , Abeygunawardana, C., Frick, D. N., Bessman, M. J., and Mildvan, A. S. (1997) Biochemistry 36, 1199-1211]. To test these proposed roles, the effects of mutations of each of these residues on the kinetic parameters and on the Mn(2+), Mg(2+), and substrate binding properties were examined. The largest decreases in k(cat) for the Mg(2+)-activated enzyme of 10(4.7)- and 10(2.6)-fold were observed for the E53Q and E53D mutants, respectively, while 97-, 48-, 25-, and 14-fold decreases were observed for the E44D, E56D, E56Q, and E44Q mutations, respectively. Smaller effects on k(cat) were observed for mutations of Glu-98 and Lys-39. For wild type MutT and its E53D and E44D mutants, plots of log(k(cat)) versus pH exhibited a limiting slope of 1 on the ascending limb and then a hump, i.e., a sharply defined maximum near pH 8 followed by a plateau, yielding apparent pK(a) values of 7.6 +/- 0.3 and 8.4 +/- 0.4 for an essential base and a nonessential acid catalyst, respectively, in the active quaternary MutT-Mg(2+)-dGTP-Mg(2+) complex. The pK(a) of 7.6 is assigned to Glu-53, functioning as a base catalyst in the active quaternary complex, on the basis of the disappearance of the ascending limb of the pH-rate profile of the E53Q mutant, and its restoration in the E53D mutant with a 10(1.9)-fold increase in (k(cat))(max). The pK(a) of 8.4 is assigned to Lys-39 on the basis of the disappearance of the descending limb of the pH-rate profile of the K39Q mutant, and the observation that removal of the positive charge of Lys-39, by either deprotonation or mutation, results in the same 8.7-fold decrease in k(cat). Values of k(cat) of both wild type MutT and the E53Q mutant were independent of solvent viscosity, indicating that a chemical step is likely to be rate-limiting with both. A liganding role for Glu-53 and Glu-56, but not Glu-98, in the binary E-M(2+) complex is indicated by the observation that the E53Q, E53D, E56Q, and E56D mutants bound Mn(2+) at the active site 36-, 27-, 4.7-, and 1.9-fold weaker, and exhibited 2.10-, 1.50-, 1.12-, and 1.24-fold lower enhanced paramagnetic effects of Mn(2+), respectively, than the wild type enzyme as detected by 1/T(1) values of water protons, consistent with the loss of a metal ligand. However, the K(m) values of Mg(2+) and Mn(2+) indicate that Glu-56, and to a lesser degree Glu-98, contribute to metal binding in the active quaternary complex. Mutations of the more distant but conserved residue Glu-44 had little effect on metal binding or enhancement factors in the binary E-M(2+) complexes. Two-dimensional (1)H-(15)N HSQC and three-dimensional (1)H-(15)N NOESY-HSQC spectra of the kinetically damaged E53Q and E56Q mutants showed largely intact proteins with structural changes near the mutated residues. Structural changes in the kinetically more damaged E44D mutant detected in (1)H-(15)N HSQC spectra were largely limited to the loop I-helix I motif, suggesting that Glu-44 stabilizes the active site region. (1)H-(15)N HSQC titrations of the E53Q, E56Q, and E44D mutants with dGTP showed changes in chemical shifts of residues lining the active site cleft, and revealed tighter nucleotide binding by these mutants, indicating an intact substrate binding site. (ABSTRACT TRUNCATED)  相似文献   

16.
We have recently described two yeast strains that are mutated in the MRF1 gene encoding the mitochondrial release factor mRF-1. Both mutants provoke gene-specific defects in mitochondrial translational termination. In the present study we report the cloning, sequencing, as well as an analysis of residual activities of both mutant mrf1 alleles. Each allele specifies a different single amino acid substitution located one amino acid apart. The amino acid changes do not affect the level or cellular localization of the mutant proteins, since equal amounts of wild type and mutant mRF-1 were detected in the mitochondrial compartment. Over-expression of the mutant alleles in wild type and mrf1 mutant yeast strains produces a phenotype consistent with a reduced affinity of the mutant release factors for the ribosome, indicating that the mutations map in a release factor domain involved in ribosome binding. We also demonstrate that nonsense suppression caused by a mutation in the mitochondrial homolog of the E. coli small ribosomal protein S4 can be reversed by a slight over-expression of the MRF1 gene.  相似文献   

17.
Hisabori T  Kondoh A  Yoshida M 《FEBS letters》1999,450(1-2):35-38
Desethylamiodarone (DEA) acts as a competitive inhibitor of triiodothyronine (T3) binding to the alpha1-thyroid hormone receptor (TR alpha1) but as a non-competitive inhibitor with respect to TR beta1. To gain insight into the position of the binding site of desethylamiodarone on TR beta1 we investigated the naturally occurring mutants Y321C, R429Q, P453A, P453T and the artificial mutants L421R and E457A in the ligand binding domain of human TR beta1. The IC50 values (in microM) of DEA for P453A (50 +/- 11) and P453T (55 +/- 16) mutant TR beta1 are not different from that for the wild type TR beta1 (56 +/- 15), but the IC50 values of R429Q (32 +/- 7; P<0.001) and E457A (17 +/- 3; P<0.001) are significantly lower than of the wild type. Scatchard plots and Langmuir analyses indicate a non-competitive nature of the inhibition by DEA of T3 binding to all four mutant TR beta1s tested. Mutants P453A and P453T do not influence overall electrostatic potential, and also do not influence the affinity for DEA compared to wild type. Mutant E457A causes a change from a negatively charged amino acid to a hydrophobic amino acid, enhancing the affinity for DEA. Mutant R429Q, located in helix 11, causes an electrostatic potential change from positive to uncharged, also resulting in greater affinity for DEA. We therefore postulate that amino acids R429 and E457 are at or close to the binding site for DEA, and that DEA does not bind in the T3 binding pocket itself, in line with the non-competitive nature of the inhibition of T3 binding to TR beta1 by DEA.  相似文献   

18.
Huang YC  Colman RF 《Biochemistry》2002,41(17):5637-5643
Sequence alignment predicts that His(309) of pig heart NADP-dependent isocitrate dehydrogenase is equivalent to His(339) of the Escherichia coli enzyme, which interacts with the coenzyme in the crystal structure [Hurley et al. (1991) Biochemistry 30, 8671-8688], and porcine His(315) and His(319) are close to that site. The mutant porcine enzymes H309Q, H309F, H315Q, and H319Q were prepared by site-directed mutagenesis, expressed in E. coli, and purified. The H319Q mutant has K(m) values for NADP, isocitrate, and Mn(2+) similar to those of wild-type enzyme, and V(max) = 20.1, as compared to 37.8 micromol of NADPH min(-1) (mg of protein)(-1) for wild type. Thus, His(319) is not involved in coenzyme binding and has a minimal effect on catalysis. In contrast, H315Q exhibits a K(m) for NADP 40 times that of wild type and V(max) = 16.2 units/mg of protein, with K(m) values for isocitrate and Mn(2+) similar to those of wild type. These results implicate His(315) in the region of the NADP site. Replacement of His(309) by Q or F yields enzyme with no detectable activity. The His(309) mutants bind NADPH poorly, under conditions in which wild type and H319Q bind 1.0 mol of NADPH/mol of subunit, indicating that His(309) is important for the binding of coenzyme. The His(309) mutants bind isocitrate stoichiometrically, as do wild-type and the other mutant enzymes. However, as distinguished from the wild-type enzyme, the His(309) mutants are not oxidatively cleaved by metal isocitrate, implying that the metal ion is not bound normally. Since circular dichroism spectra are similar for wild type, H315Q, and H319Q, these amino acid substitutions do not cause major conformational changes. In contrast, replacement of His(309) results in detectable change in the enzyme's CD spectrum and therefore in its secondary structure. We propose that His(309) plays a significant role in the binding of coenzyme, contributes to the proper coordination of divalent metal ion in the presence of isocitrate, and maintains the normal conformation of the enzyme.  相似文献   

19.
20.
De novo mutations in ATP1A3, the gene encoding the α3-subunit of Na+,K+-ATPase, are associated with the neurodevelopmental disorder Alternating Hemiplegia of Childhood (AHC). The aim of this study was to determine the functional consequences of six ATP1A3 mutations (S137Y, D220N, I274N, D801N, E815K, and G947R) associated with AHC. Wild type and mutant Na+,K+-ATPases were expressed in Sf9 insect cells using the baculovirus expression system. Ouabain binding, ATPase activity, and phosphorylation were absent in mutants I274N, E815K and G947R. Mutants S137Y and D801N were able to bind ouabain, although these mutants lacked ATPase activity, phosphorylation, and the K+/ouabain antagonism indicative of modifications in the cation binding site. Mutant D220N showed similar ouabain binding, ATPase activity, and phosphorylation to wild type Na+,K+-ATPase. Functional impairment of Na+,K+-ATPase in mutants S137Y, I274N, D801N, E815K, and G947R might explain why patients having these mutations suffer from AHC. Moreover, mutant D801N is able to bind ouabain, whereas mutant E815K shows a complete loss of function, possibly explaining the different phenotypes for these mutations.  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