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.
Reduction potential determinations of K75E, E139K and E301A ferredoxin-NADP+ reductases provide valuable information concerning the factors that contribute to tune the flavin reduction potential. Thus, while E139 is not involved in such modulation, the K75 side-chain tunes the flavin potential by creating a defined environment that modulates the FAD conformation. Finally, the E301 side-chain influences not only the flavin reduction potential, but also the electron transfer mechanism, as suggested from the values determined for the E301A mutant, where E(ox/rd) and E(sq/rd) shifted +41 and +102 mV, respectively, with regard to wild-type. Reduction potentials allowed estimation of binding energies differences of the FAD cofactor upon reduction.  相似文献   

2.
The electron transfer cascade from photosystem I to NADP+ was studied at physiological pH by flash-absorption spectroscopy in a Synechocystis PCC6803 reconstituted system comprised of purified photosystem I, ferredoxin, and ferredoxin-NADP+ reductase. Experiments were conducted with a 34-kDa ferredoxin-NADP+ reductase homologous to the chloroplast enzyme and a 38-kDa N-terminal extended form. Small differences in kinetic and catalytic properties were found for these two forms, although the largest one has a 3-fold decreased affinity for ferredoxin. The dissociation rate of reduced ferredoxin from photosystem I (800 s(-1)) and the redox potential of the first reduction of ferredoxin-NADP+ reductase (-380 mV) were determined. In the absence of NADP+, differential absorption spectra support the existence of a high affinity complex between oxidized ferredoxin and semireduced ferredoxin-NADP+ reductase. An effective rate of 140-170 s(-1) was also measured for the second reduction of ferredoxin-NADP+ reductase, this process having a rate constant similar to that of the first reduction. In the presence of NADP+, the second-order rate constant for the first reduction of ferredoxin-NADP+ reductase was 20% slower than in its absence, in line with the existence of ternary complexes (ferredoxin-NADP+ reductase)-NADP+-ferredoxin. A single catalytic turnover was monitored, with 50% NADP+ being reduced in 8-10 ms using 1.6 microM photosystem I. In conditions of multiple turnover, we determined initial rates of 360-410 electrons per s and per ferredox-in-NADP+ reductase for the reoxidation of 3.5 microM photoreduced ferredoxin. Identical rates were found with photosystem I lacking the PsaE subunit and wild type photosystem I. This suggests that, in contrast with previous proposals, the PsaE subunit is not involved in NADP+ photoreduction.  相似文献   

3.
The redox-induced structural changes at the active site of the superoxide reductase (SOR) from Desulfoarculus baarsii and Treponema pallidum have been monitored by means of FTIR difference spectroscopy coupled to electrochemistry. With this technique, the structure and interactions formed by individual amino acids at a redox site can be detected. The infrared data on wild-type, Glu47Ala, and Lys48Ile mutants of the SOR from D. baarsii provide experimental support for the conclusion that the two different coordination motifs observed in the three-dimensional structure of the SOR from Pyrococcus furiosus [Yeh, A. P., Hu, Y., Jenney, F. E., Adams, M. W. W., and Rees, D. (2000) Biochemistry 39, 2499-2508] correspond to the two redox forms of the SOR iron center. We extend this result to the center II iron of SOR of the desulfoferrodoxin type. Similar structural changes are also observed upon iron oxidation in the SOR of T. pallidum. In D. baarsii, the IR modes of the Glu47 side chain support that it provides a monodentate ligand to the oxidized iron, while it does not interact with Fe(2+). Structural changes at the level of peptide bond(s) observed upon iron oxidation in wild-type are suppressed in the Glu47Ala mutant. We propose that the presence of the Glu side chain plays an important role for the structural reorganization accompanying iron oxidation. We identified the infrared modes of the Lys48 side chain and found that a change in its environment occurs upon iron oxidation. The lack of other structural changes upon the Lys48Ile mutation shows that the catalytic role of Lys, as evidenced by pulse radiolysis experiments [Lombard, M., Houée-Levin, C., Touati, D., Fontecave, M., and Nivière, V. (2001) Biochemistry 40, 5032-5040], is purely electrostatic, guiding superoxide toward the reduced iron.  相似文献   

4.
Kim TJ  Park CS  Cho HY  Cha SS  Kim JS  Lee SB  Moon TW  Kim JW  Oh BH  Park KH 《Biochemistry》2000,39(23):6773-6780
A sequence alignment shows that residue 332 is conserved as glutamate in maltogenic amylases (MAases) and in other related enzymes such as cyclodextrinase and neopullulanase, while the corresponding position is conserved as histidine in alpha-amylases. We analyzed the role of Glu332 in the hydrolysis and the transglycosylation activity of Thermus MAase (ThMA) by site-directed mutagenesis. Replacing Glu332 with histidine reduced transglycosylation activity significantly, but enhanced hydrolysis activity on alpha-(1,3)-, alpha-(1,4)-, and alpha-(1,6)-glycosidic bonds relative to the wild-type (WT) enzyme. The mutant Glu332Asp had catalytic properties similar to those of the WT enzyme, but the mutant Glu332Gln resulted in significantly decreased transglycosylation activity. These results suggest that an acidic side chain at position 332 of MAase plays an important role in the formation and accumulation of transfer products by modulating the relative rates of hydrolysis and transglycosylation. From the structure, we propose that an acidic side chain at position 332, which is located in a pocket, is involved in aligning the acceptor molecule to compete with water molecules in the nucleophilic attack of the glycosyl-enzyme intermediate.  相似文献   

5.
Mao Y  Ba Y 《Biophysical journal》2006,91(3):1059-1068
The primary sequences of type I antifreeze proteins (AFPs) are Ala rich and contain three 11-residue repeat units beginning with threonine residues. Their secondary structures consist of alpha-helices. Previous activity study of side-chain mutated AFPs suggests that the ice-binding side of type I AFPs comprises the Thr side chains and the conserved i + 4 and i + 8 Ala residues, where i indicates the positions of the Thrs. To find structural evidence for the AFP's ice-binding side, a variable-temperature dependent (13)C spin lattice relaxation solid-state NMR experiment was carried out for two Ala side chain (13)C labeled HPLC6 isoforms of the type I AFPs each frozen in H(2)O and D(2)O, respectively. The first one was labeled on the equivalent 17th and 21st Ala side chains (i + 4, 8), and the second one on the equivalent 8th, 19th, and 30th Ala side chains (i + 6). The two kinds of labels are on the opposite sides of the alpha-helical AFP. A model of Ala methyl group rotation/three-site rotational jump combined with water molecular reorientation was tested to probe the interactions of the methyl groups with the proximate water molecules. Analysis of the T(1) data shows that there could be 10 water molecules closely capping an i + 4 or an i + 8 methyl group within the range of van der Waals interaction, whereas the surrounding water molecules to the i + 6 methyl groups could be looser. This study suggests that the side of the alpha-helical AFP comprising the i + 4 and i + 8 Ala methyl groups could interact with the ice surface in the ice/water interface.  相似文献   

6.
Two mutants of the spinach ferredoxin-NADP+ reductase (FNR) were constructed, expressed by using a heterologous expression system previously described (Aliverti, A., Jansen, T., Zanetti, G., Ronchi, S., Herrmann, R. G., and Curti, B. (1990) Eur. J. Biochem. 191, 551-555), and purified to homogeneity. The mutant enzymes FNR-Lys116Gln and FNR-Lys244Gln were similar to the wild-type enzyme in the physicochemical properties measured; however, both enzymes showed a lower activity. Steady-state kinetic analyses and NADP+ binding measurements of the mutant proteins have revealed that the Lys116Gln enzyme had a catalytic efficiency (kcat/KmNADPH) 500-fold lower than that of the wild-type enzyme, and its interaction with NADP+ was greatly impaired. The Lys244Gln enzyme instead had almost the same kcat in the ferricyanide reductase reaction as the wild-type enzyme, while higher values of KmNADPH and Kd for NADP+ were observed. Thus, protein engineering has enabled us to define the role of Lys-116 and Lys-244 in ferredoxin-NADP+ reductase; whereas Lys-244 contributes to the stabilization of the Michaelis complex, the Lys-116 side chain plays an important role also in the further steps of the catalytic cycle, because both Km and kcat are highly affected by the mutation.  相似文献   

7.
M D Distefano  K G Au  C T Walsh 《Biochemistry》1989,28(3):1168-1183
Mercuric reductase, a flavoenzyme that possess a redox-active cystine, Cys135Cys140, catalyzes the reduction of Hg(II) to Hg(0) by NADPH. As a probe of mechanism, we have constructed mutants lacking a redox-active disulfide by eliminating Cys135 (Ala135Cys140), Cys140 (Cys135Ala140), or both (Ala135Ala140). Additionally, we have made double mutants that lack Cys135 (Ala135Cys139Cys140) or Cys140 (Cys135Cys139Ala140) but introduce a new Cys in place of Gly139 with the aim of constructing dithiol pairs in the active site that do not form a redox-active disulfide. The resulting mutant enzymes all lack redox-active disulfides and are hence restricted to FAD/FADH2 redox chemistry. Each mutant enzyme possesses unique physical and spectroscopic properties that reflect subtle differences in the FAD microenvironment. These differences are manifested in a 23-nm range in enzyme-bound FAD lambda max values, an 80-nm range in thiolate to flavin charge-transfer absorbance maxima, and a ca. 100-mV range in FAD reduction potential. Preliminary evidence for the Ala135Cys139Cys140 mutant enzyme suggests that this protein forms a disulfide between the two adjacent Cys residues. Hg(II) titration experiments that correlate the extent of charge-transfer quenching with Hg(II) binding indicate that the Ala135Cys140 protein binds Hg(II) with substantially less avidity than does the wild-type enzyme. All mutant mercuric reductases catalyze transhydrogenation and oxygen reduction reactions through obligatory reduced flavin intermediates at rates comparable to or greater than that of the wild-type enzyme. For these activities, there is a linear correlation between log kappa cat and enzyme-bound FAD reduction potential. In a sensitive Hg(II)-mediated enzyme-bound FADH2 reoxidation assay, all mutant enzymes were able to undergo at least one catalytic event at rates 50-1000-fold slower than that of the wild-type enzyme. We have also observed the reduction of Hg(II) by free FADH2. In multiple-turnover assays which monitored the production of Hg(0), two of the mutant enzymes were observed to proceed through at least 30 turnovers at rates ca. 1000-fold slower than that of wild-type mercuric reductase. We conclude that the Cys135 and Cys140 thiols serve as Hg(II) ligands that orient the Hg(II) for subsequent reduction by a reduced flavin intermediate.  相似文献   

8.
Src Homology (SH2) domains play critical roles in signaling pathways by binding to phosphotyrosine (pTyr)-containing sequences, thereby recruiting SH2 domain-containing proteins to tyrosine-phosphorylated sites on receptor molecules. Investigations of the peptide binding specificity of the SH2 domain of the Src kinase (Src SH2 domain) have defined the EEI motif C-terminal to the phosphotyrosine as the preferential binding sequence. A subsequent study that probed the importance of eight specificity-determining residues of the Src SH2 domain found two residues which when mutated to Ala had significant effects on binding: Tyr beta D5 and Lys beta D3. The mutation of Lys beta D3 to Ala was particularly intriguing, since a Glu to Ala mutation at the first (+1) position of the EEI motif (the residue interacting with Lys beta D3) did not significantly affect binding. Hence, the interaction between Lys beta D3 and +1 Glu is energetically coupled. This study is focused on the dissection of the energetic coupling observed across the SH2 domain-phosphopeptide interface at and around the +1 position of the peptide. It was found that three residues of the SH2 domain, Lys beta D3, Asp beta C8 and AspCD2 (altogether forming the so-called +1 binding region) contribute to the selection of Glu at the +1 position of the ligand. A double (Asp beta C8Ala, AspCD2Ala) mutant does not exhibit energetic coupling between Lys beta D3 and +1 Glu, and binds to the pYEEI sequence 0.3 kcal/mol tighter than the wild-type Src SH2 domain. These results suggest that Lys beta D3 in the double mutant is now free to interact with the +1 Glu and that the role of Lys beta D3 in the wild-type is to neutralize the acidic patch formed by Asp beta C8 and AspCD2 rather than specifically select for a Glu at the +1 position as it had been hypothesized previously. A triple mutant (Lys beta D3Ala, Asp beta C8Ala, AspCD2Ala) has reduced binding affinity compared to the double (Asp beta C8Ala, AspCD2Ala) mutant, yet binds the pYEEI peptide as well as the wild-type Src SH2 domain. The structural basis for such high affinity interaction was investigated crystallographically by determining the structure of the triple (Lys beta D3Ala, Asp beta C8Ala, AspCD2Ala) mutant bound to the octapeptide PQpYEEIPI (where pY indicates a phosphotyrosine). This structure reveals for the first time contacts between the SH2 domain and the -1 and -2 positions of the peptide (i.e. the two residues N-terminal to pY). Thus, unexpectedly, mutations in the +1 binding region affect binding of other regions of the peptide. Such additional contacts may account for the high affinity interaction of the triple mutant for the pYEEI-containing peptide.  相似文献   

9.
Flavoenzymes may reduce quinones in a single-electron, mixed single- and two-electron, and two-electron way. The mechanisms of two-electron reduction of quinones are insufficiently understood. To get an insight into the role of flavin semiquinone stability in the regulation of single- vs. two-electron reduction of quinones, we studied the reactions of wild type Anabaena ferredoxin:NADP(+)reductase (FNR) with 48% FAD semiquinone (FADH*) stabilized at the equilibrium (pH 7.0), and its Glu301Ala mutant (8% FADH* at the equilibrium). We found that Glu301Ala substitution does not change the quinone substrate specificity of FNR. However, it confers the mixed single- and two-electron mechanism of quinone reduction (50% single-electron flux), whereas the wild type FNR reduces quinones in a single-electron way. During the oxidation of fully reduced wild type FNR by tetramethyl-1,4-benzoquinone, the first electron transfer (formation of FADH*) is about 40 times faster than the second one (oxidation of FADH*). In contrast, the first and second electron transfer proceeded at similar rates in Glu301Ala FNR. Thus, the change in the quinone reduction mechanism may be explained by the relative increase in the rate of second electron transfer. This enabled us to propose the unified scheme of single-, two- and mixed single- and two-electron reduction of quinones by flavoenzymes with the central role of the stability of flavin/quinone ion-radical pair.  相似文献   

10.
Tu X  Hubbard PA  Kim JJ  Schulz H 《Biochemistry》2008,47(4):1167-1175
NADPH-dependent 2,4-dienoyl-CoA reductase (DCR) is one of the auxiliary enzymes required for the beta-oxidation of unsaturated fatty acids. Mutants of Escherichia coli DCR were generated by site-directed mutagenesis to explore the molecular mechanism of this enzyme. The Tyr166Phe mutant, which was expected to be inactive due to the loss of its putative proton donor residue, exhibited 27% of the wild-type activity. However, the product of the reduction was 3-enoyl-CoA instead of 2-enoyl-CoA, the normal product. Glu164 seems to function as proton donor in the Tyr166Phe mutant, because the Tyr166Phe/ Glu164Gln double mutant was inactive whereas the Glu164Ala mutant exhibited low but significant activity. His252 is important for the efficient operation of Tyr166 because a His252Ala mutation by itself reduced the activity of DCR by 3 orders of magnitude, whereas the Tyr166Phe/His252Ala double mutation exhibited 4.4% of the wild-type activity. This data supports a mechanism that has Tyr166 with the assistance of His252 acting as proton donor in the wild-type enzyme to produce 2-enoyl-CoA, whereas Glu164 serves as the proton donor in the absence of Tyr166 to yield 3-enoyl-CoA. A Cys337Ala mutation, which resulted in the loss of most of the iron and acid-labile sulfur, decreased the reductase activity more than 1000-fold. This observation agrees with the proposed operation of an intramolecular electron transport chain that is essential for the effective catalysis of E. coli DCR.  相似文献   

11.
The structure and self‐assembly of the peptide corresponding to the third transmembrane domain (TMD3) of Slc11a1 and its E139A mutant are studied in 1,1,1,3,3,3‐hexafluoro‐2‐propanol (HFIP) aqueous solution by NMR and CD experiments. Slc11a1 is an integral membrane protein with 12 putative TMDs and functions as a pH‐coupled divalent metal cation transporter. Glu139 of Slc11a1 is highly conserved within predicted TMD3 of the Slc11 protein family and function‐associated. Here, we provide the first direct experimental evidence for the structural features of two 24‐residue peptides corresponding to TMD3 of Slc11a1 and its E139A mutant in 60% HFIP‐d2 aqueous solution using CD and NMR spectroscopies. Our study shows that the membrane‐spanning peptide folds as a typical amphipathic α‐helix structure from Ile5 to Met20 with hydrophilic residues Glu12 (Glu139 in Slc11a1) and Asp19 lying on the same side of the helix. The substitution of Glu139 by an alanine residue has little effect on the structure of the peptide, but increases hydrophobicity and facilitates self‐assembly of the peptide. Although the wildtype peptide is monomeric in HFIP aqueous solution, the E139A mutant forms a dimer. The increase in hydrophobicity of the membrane‐spanning peptide and/or change in the interactions between transmembrane segments induced by E139A mutation may affect the metal ion transport of the protein. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
Polar mutations in transmembrane alpha helices may alter the structural details of the hydrophobic sequences and control intermolecular contacts. We have performed molecular dynamics simulations on the transmembrane domain of the proto-oncogenic and the oncogenic forms of the Neu receptor in a fluid DMPC bilayer to test whether the Glu mutation which replaces the Val residue at position 664 may alter the helical structure and its insertion in the membrane. The simulations show that the wild and the mutant forms of the transmembrane domain have a different behavior in the bilayer. The native transmembrane sequence is found to be more flexible than in the presence of the Glu mutation, characterized by a tendency to pi deformation to accommodate the helix length to the membrane thickness. The mutant form of this domain does not evidence helical deformation in the present simulation. Hydrophobic matching is achieved both by a larger helix tilt and a vertical shift of the helix towards the membrane interface, favoring the accessibility of the Glu side chain to the membrane environment. A rapid exchange of hydrogen bond interactions with the surrounding water molecules and the lipid headgroups is observed. The difference in the behavior between the two peptides in a membrane environment was also observed experimentally. Both simulation and experimental results agree with the hypothesis that water may act as an intermediate for the formation of cross links between the facing Glu side chains stabilizing the dimer.  相似文献   

13.
Five moeA mutants were generated by replacing some conserved amino acids of MoeA by site-directed mutagenesis. The mutants were assayed for the ability to restore in vivo nitrate reductase activity of the moeA mutant Escherichia coli JRG97 and in vitro Neurospora crassa nit-1 nitrate reductase activity. The replacements Asp59AlaGly60Ala, Asp259Ala, Pro298AlaPro301Ala abolished the function of MoeA in Mo-molybdopterin formation and stabilization, reflected in the inability to restore nitrate reductase activity. The replacements Gly251AlaGly252Ala reduced, and that of Pro283Ala had no effect, on nitrate reductase activity. E. coli JRG97 cells transformed with mutants that failed to restore nitrate reductase activity showed by HPLC analysis a decreased level of molybdopterin-derived dephospho FormA as compared to bacteria transformed with wild-type moeA. The effects of the amino acid replacements on MoeA function may be explained in correlation with the MoeA crystal structure.  相似文献   

14.
Pseudoazurin (a blue copper protein or cupredoxin) of a denitrifying bacterium Alcaligenes faecalis S-6 is a direct electron carrier for a Cu-containing nitrite reductase (NIR) of the same organism. Site-directed mutagenesis of the pseudoazurin was carried out using an Escherichia coli expression system. Replacement of Tyr74 by Phe to remove an internal hydrogen bond in the beta-barrel caused a slight decrease in heat stability as well as a requirement for a higher concentration of Cu2+ for production in the E. coli host. Exchange of Ala for Pro80 adjacent to His81, one of the four ligands binding a type I Cu atom, caused a marked increase in reduction potential by 139 mV without change in the optical absorption spectrum. The ability of the pseudoazurin to transfer electrons to NIR was markedly diminished but the apparent Km of NIR for pseudoazurin was not affected by the mutation. X-ray diffraction data collected on the oxidized and reduced forms of the Pro80Ala mutant show that a water molecule occupies the pocket created by the absent side chain. This observation suggests that the increase in reduction potential may be caused due to the increased solvent accessibility to the Cu atom. The electron density difference maps on these structures (at 2.0 A) show that this water moves during the change in oxidation state, and that there are small, but localized, conformational changes greater than 6.5 A from the copper site, as well as movement of both the Cu2+ and the cysteinate sulfur.  相似文献   

15.
The phenolic side chain of Tyr(4) present in Ang II is proposed to interact with the side chain of Arg 167 of the AT1 receptor. To determine the contribution of the analogous Arg182 in the ligand-binding properties of the AT2, we replaced the Arg182 with Glu and Ala, and analyzed the ligand-binding properties. Our results suggest that replacing Arg182 with either Glu or Ala abolished the ability of the AT2 receptor to bind the nonspecific peptidic ligands, (125)I-Ang II and [(125)I-Sar(1)-Ile(8)]Ang II, as well as the AT2 receptor-specific peptidic ligand (125)I-CGP42112A. We have shown previously that replacing the positively charged side chain of Lys215 with the negatively charged side chain of Glu in the fifth TMD did not alter the high affinity binding of (125)I-CGP42112A to the AT2 receptor. However, ligand-binding properties of the Arg182Glu mutant suggest that positively charged side chain of Arg182 located in the junction of second ECL and the fourth TMD is critical for high affinity binding of all three peptidic ligands to the AT2 receptor.  相似文献   

16.
T D Porter  C B Kasper 《Biochemistry》1986,25(7):1682-1687
The FMN-binding domain of NADPH-cytochrome P-450 oxidoreductase, residues 77-228, is homologous with bacterial flavodoxins, while the FAD-binding domain, residues 267-678, shows a high degree of similarity to two FAD-containing proteins, ferredoxin-NADP+ reductase and NADH-cytochrome b5 reductase. Comparison of these proteins to glutathione reductase, a flavoprotein whose three-dimensional structure is known, has permitted tentative identification of FAD- and cofactor-binding residues in these proteins. The remarkable conservation of sequence between NADPH-cytochrome P-450 oxidoreductase and ferredoxin-NADP+ reductase, coupled with the homology of the FMN-binding domain of the oxidoreductase with the bacterial flavodoxins, implies that NADPH-cytochrome P-450 oxidoreductase arose as a result of fusion of the ancestral genes for these two functionally linked flavoproteins.  相似文献   

17.
Hermes S  Stachnik JM  Onidas D  Remy A  Hofmann E  Gerwert K 《Biochemistry》2006,45(46):13741-13749
The reaction center (RC) of Rhodobacter sphaeroides uses light energy to reduce and protonate a quinone molecule, QB (the secondary quinone electron acceptor), to form quinol, QBH2. Asp210 in the L-subunit has been shown to be a catalytic residue in this process. Mutation of Asp210 to Asn leads to a deceleration of reoxidation of QA- in the QA-QB --> QAQB- transition. Here we determined the structure of the Asp210 to Asn mutant to 2.5 A and show that there are no major structural differences as compared to the wild-type protein. We found QB in the distal position and a chain of water molecules between Asn210 and QB. Using time-resolved Fourier transform infrared (trFTIR) spectroscopy, we characterized the molecular reaction mechanism of this mutant. We found that QB- formation precedes QA- oxidation even more pronounced than in the wild-type reaction center. Continuum absorbance changes indicate deprotonation of a protonated water cluster, most likely of the water chain between Asn210 and QB. A detailed analysis of wild-type structures revealed a highly conserved water chain between Asp210 or Glu210 and QB in Rb. sphaeroides and Rhodopseudomonas viridis, respectively.  相似文献   

18.
Structural and enzymological studies have shown the importance of Glu144 and Glu164 for the catalysis by 2-enoyl-CoA hydratase-1 (crotonase). Here we report about the enzymological properties of the Glu144Ala and Glu164Ala variants of rat mitochondrial 2-enoyl-CoA hydratase-1. Size-exclusion chromatography and CD spectroscopy showed that the wild-type protein and mutants have similar oligomerization states and folding. The kcat values of the active site mutants Glu144Ala and Glu164Ala were decreased about 2000-fold, but the Km values were unchanged. For study of the potential intrinsic Delta3-Delta2-enoyl-CoA isomerase activity of mECH-1, a new assay using 2-enoyl-CoA hydratase-2 and (R)-3-hydroxyacyl-CoA dehydrogenase as auxiliary enzymes was introduced. It was demonstrated that rat wild-type mECH-1 is also capable of catalyzing isomerization with the activity ratio (isomerization/hydration) of 1/5000. The kcat values of isomerization in Glu144Ala and Glu164Ala were decreased 10-fold and 1000-fold, respectively. The data are in line with the proposal that Glu164 acts as a protic amino acid residue for both the hydration and the isomerization reaction. The structural factors favoring the hydratase over the isomerase reaction have been addressed by investigating the enzymological properties of the Gln162Ala, Gln162Met, and Gln162Leu variants. The Gln162 side chain is hydrogen bonded to the Glu164 side chain; nevertheless, these mutants have enzymatic properties similar to that of the wild type, indicating that catalytic function of the Glu164 side chain in the hydratase and isomerase reaction does not depend on the interactions with the Gln162 side chain.  相似文献   

19.
Cytochrome c oxidase (CytcO) is a redox-driven, membrane-bound proton pump. One of the proton transfer pathways of the enzyme, the D pathway, used for the transfer of both substrate and pumped protons, accommodates a network of hydrogen-bonded water molecules that span the distance between an aspartate (Asp(132)), near the protein surface, and glutamate Glu(286), which is an internal proton donor to the catalytic site. To investigate how changes in the environment around Glu(286) affect the mechanism of proton transfer through the pathway, we introduced a non-hydrogen-bonding (Ala) or an acidic residue (Asp) at position Ser(197) (S197A or S197D), located approximately 7 A from Glu(286). Although Ser(197) is hydrogen-bonded to a water molecule that is part of the D pathway "proton wire," replacement of the Ser by an Ala did not affect the proton transfer rate. In contrast, the S197D mutant CytcO displayed a turnover activity of approximately 35% of that of the wild-type CytcO, and the O(2) reduction reaction was not linked to proton pumping. Instead, a fraction of the substrate protons was taken from the positive ("incorrect") side of the membrane. Furthermore, the pH dependence of the proton transfer rate was altered in the mutant CytcO. The results indicate that there is plasticity in the water coordination of the proton pathway, but alteration of the electrostatic potential within the pathway results in uncoupling of the proton translocation machinery.  相似文献   

20.
Protein folding problem remains a formidable challenge as main chain, side chain and solvent interactions remain entangled and have been difficult to resolve. Alanine‐based short peptides are promising models to dissect protein folding initiation and propagation structurally as well as energetically. The effect of N‐terminal diproline and charged side chains is assessed on the stabilization of helical conformation in alanine‐based short peptides using circular dichroism (CD) with water and methanol as solvent. A1 (Ac–Pro–Pro–Ala–Lys–Ala–Lys–Ala–Lys–Ala–NH2) is designed to assess the effect of N‐terminal homochiral diproline and lysine side chains to induce helical conformation. A2 (Ac–Pro–Pro–Glu–Glu–Ala–Ala–Lys–Lys–Ala–NH2) and A3 (Ac–d Pro–Pro–Glu–Glu–Ala–Ala–Lys–Lys–Ala–NH2) with N‐terminal homochiral and heterochiral diproline, respectively, are designed to assess the effect of Glu...Lys (i , i  + 4) salt bridge interactions on the stabilization of helical conformation. The CD spectra of A1 , A2 and A3 in water manifest different amplitudes of the observed polyproline II (PPII) signals, which indicate different conformational distributions of the polypeptide structure. The strong effect of solvent substitution from water to methanol is observed for the peptides, and CD spectra in methanol evidence A2 and A3 as helical folds. Temperature‐dependent CD spectra of A1 and A2 in water depict an isodichroic point reflecting coexistence of two conformations, PPII and β‐strand conformation, which is consistent with the previous studies. The results illuminate the effect of N‐terminal diproline and charged side chains in dictating the preferences for extended‐β, semi‐extended PPII and helical conformation in alanine‐based short peptides. The results of the present study will enhance our understanding on stabilization of helical conformation in short peptides and hence aid in the design of novel peptides with helical structures. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  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