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.
E E Howell  C Booth  M Farnum  J Kraut  M S Warren 《Biochemistry》1990,29(37):8561-8569
The adaptability of Escherichia coli dihydrofolate reductase (DHFR) is being explored by identifying second-site mutations that can partially suppress the deleterious effect associated with removal of the active-site proton donor aspartic acid-27. The Asp27----serine mutant DHFR (D27S) was previously characterized and the catalytic activity found to be greatly decreased at pH 7.0 [Howell et al. (1986) Science 231, 1123-1128]. Using resistance to trimethoprim (a DHFR inhibitor) in a genetic selection procedure, we have isolated a double-mutant DHFR gene containing Asp27----Ser and Phe137----Ser mutations (D27S+F137S). The presence of the F137S mutation increases kcat approximately 3-fold and decreases Km(DHF) approximately 2-fold over D27S DHFR values. The overall effect on kcat/Km(DHF) is a 7-fold increase. The D27S+F137S double-mutant DHFR is still 500-fold less active than wild-type DHFR at pH 7. Surprisingly, Phe137 is approximately 15 A from residue 27 in the active site and is part of a beta-bulge. We propose the F137S mutation likely causes its catalytic effect by slightly altering the conformation of D27S DHFR. This supposition is supported by the observation that the F137S mutation does not have the same kinetic effect when introduced into the wild-type and D27S DHFRs, by the altered distribution of two conformers of free enzyme [see Dunn et al. (1990)] and by a preliminary difference Fourier map comparing the D27S and D27S+F137S DHFR crystal structures.  相似文献   

2.
During the initial stages of the bacteriorhodopsin photocycle, a proton is transferred from the Schiff base to the deprotonated carboxylate of Asp85. Earlier studies have shown that replacement of Asp85 by Asn completely abolishes proton transport activity, whereas extension of the side chain by an additional carbon-carbon bond (Asp85-->Glu) results in a functional proton pump. Here we show that extension of the Asp85 side chain by two additional bond lengths also results in a functional proton pump as long as the terminal group is a carboxylate moiety. These side chains were created by modification of the cysteine residue in the Asp85-->Cys mutant with either iodoacetic acid or iodoacetamide. In vitro chromophore formation studies show that the rate of Schiff base protonation in mutants that contain a carboxylate at residue 85 is invariably faster than in mutants that contain neutral substitutions at this position. We conclude that in bacteriorhodopsin, there is considerable tolerance in the volume of the side chain that can be accommodated at position 85 and that the presence of a carboxylate at residue 85 is important both for proton pumping and for stabilizing the protonated Schiff base.  相似文献   

3.
We have applied site-directed mutagenesis methods to change the conserved tryptophan-22 in the substrate binding site of Escherichia coli dihydrofolate reductase to phenylalanine (W22F) and histidine (W22H). The crystal structure of the W22F mutant in a binary complex with the inhibitor methotrexate has been refined at 1.9-A resolution. The W22F difference Fourier map and least-squares refinement show that structural effects of the mutation are confined to the immediate vicinity of position 22 and include an unanticipated 0.4-A movement of the methionine-20 side chain. A conserved bound water-403, suspected to play a role in the protonation of substrate DHF, has not been displaced by the mutation despite the loss of a hydrogen bond with tryptophan-22. Steady-state kinetics, stopped-flow kinetics, and primary isotope effects indicate that both mutations increase the rate of product tetrahydrofolate release, the rate-limiting step in the case of the wild-type enzyme, while slowing the rate of hydride transfer to the point where it now becomes at least partially rate determining. Steady-state kinetics show that below pH 6.8, kcat is elevated by up to 5-fold in the W22F mutant as compared with the wild-type enzyme, although kcat/Km(dihydrofolate) is lower throughout the observed pH range. For the W22H mutant, both kcat and kcat/Km(dihydrofolate) are substantially lower than the corresponding wild-type values. While both mutations weaken dihydrofolate binding, cofactor NADPH binding is not significantly altered. Fitting of the kinetic pH profiles to a general protonation scheme suggests that the proton affinity of dihydrofolate may be enhanced upon binding to the enzyme. We suggest that the function of tryptophan-22 may be to properly position the side chain of methionine-20 with respect to N5 of the substrate dihydrofolate.  相似文献   

4.
In ferredoxin I from Azotobacter vinelandii, the reduction of a [3Fe-4S] iron-sulphur cluster is coupled with the protonation of the mu2S sulphur atom that is approx. 6 A away from the protein boundary. The recent study of the site-specific mutants of ferredoxin I led to the conclusion that a particular surface aspartic residue (Asp15) is solely responsible for the proton transfer to the mu2S atom by 'rapid penetrative excursions' (K. Chen, J. Hirst, R. Camba, C.A. Bonagura, C.D. Stout, B.K. Burgess, F.A. Armstrong, Nature 405 (2000) 814-817). In the same paper it has been reported that the replacement of Asp15 by glutamate led to the blockage of the enzyme, although glutamate, with its longer and more flexible side chain, should apparently do even better as a mobile proton carrier than aspartate. We tackled this puzzling incompetence of Glu15 by molecular dynamics simulations. It was revealed that the conformational alterations of Asp15 are energetically balanced by the straining of the nearby Lys84 side chain in wild-type ferredoxin I but not in the Asp15-->Glu mutant. Lys84 in ferredoxin I of A. vinelandii seems to represent the first case where the strained (entatic) conformation of a particular amino acid side chain could be directly identified in the ground state of an enzyme and assigned to a distinct mechanism of energy balance during the catalytic transition.  相似文献   

5.
We have explored the substrate protonation mechanism of Escherichia coli dihydrofolate reductase by changing the location of the proton donor. A double mutant was constructed in which the proton donor of the wild-type enzyme, aspartic acid-27, has been changed to serine and simultaneously an alternative proton donor, glutamic acid, has replaced threonine at position 113. The active site of the resulting variant enzyme molecule should therefore somewhat resemble that proposed for the R67 plasmid-encoded dihydrofolate reductase [Matthews, D. A., Smith, S. L., Baccanari, D. P., Burchall, J. J., Oatley, S. J., & Kraut, J. (1986) Biochemistry 25, 4194]. At pH 7, the double-mutant enzyme has a 3-fold greater kcat and an unchanged Km(dihydrofolate) as compared with the single-mutant Asp-27----Ser enzyme described previously [Howell, E. E., Villafranca, J. E., Warren, M. S., Oatley, S. J., & Kraut, J. (1986) Science (Washington, D.C.) 231, 1123]. Additionally, its activity vs pH profiles together with observed deuterium isotope effects, suggest that catalysis depends on an acidic group with a pKa of 8. It is concluded that the dihydropteridine ring of a bound substrate molecule can indeed be protonated by a glutamic acid side chain at position 113 (instead of an aspartic acid side chain at position 27), but with greatly decreased efficiency: at pH 7, the double mutant still has a 25-fold lower kcat (1.2 s-1) and a 2900-fold lower kcat/km(dihydrofolate) (8.6 X 10(3) s-1 M-1) than the wild-type enzyme.  相似文献   

6.
To understand the pressure-adaptation mechanism of deep-sea enzymes, we studied the effects of pressure on the enzyme activity and structural stability of dihydrofolate reductase (DHFR) of the deep-sea bacterium Moritella profunda (mpDHFR) in comparison with those of Escherichia coli (ecDHFR). mpDHFR exhibited optimal enzyme activity at 50MPa whereas ecDHFR was monotonically inactivated by pressure, suggesting inherent pressure-adaptation mechanisms in mpDHFR. The secondary structure of apo-mpDHFR was stable up to 80°C, as revealed by circular dichroism spectra. The free energy changes due to pressure and urea unfolding of apo-mpDHFR, determined by fluorescence spectroscopy, were smaller than those of ecDHFR, indicating the unstable structure of mpDHFR against pressure and urea despite the three-dimensional crystal structures of both DHFRs being almost the same. The respective volume changes due to pressure and urea unfolding were -45 and -53ml/mol at 25°C for mpDHFR, which were smaller (less negative) than the corresponding values of -77 and -85ml/mol for ecDHFR. These volume changes can be ascribed to the difference in internal cavity and surface hydration of each DHFR. From these results, we assume that the native structure of mpDHFR is loosely packed and highly hydrated compared with that of ecDHFR in solution.  相似文献   

7.
The active-site cysteines (Cys 32 and Cys 35) of Escherichia coli thioredoxin are oxidized to a disulfide bridge when the protein mediates substrate reduction. In reduced thioredoxin, Cys 32 and Cys 35 are characterized by abnormally low pKa values. A conserved side chain, Asp 26, which is sterically accessible to the active site, is also essential to oxidoreductase activity. pKa values governing cysteine thiol-thiolate equilibria in the mutant thioredoxin, D26A, have been determined by direct Raman spectrophotometric measurement of sulfhydryl ionizations. The results indicate that, in D26A thioredoxin, both sulfhydryls titrate with apparent pKa values of 7.5+/-0.2, close to values measured previously for wild-type thioredoxin. Sulfhydryl Raman markers of D26A and wild-type thioredoxin also exhibit similar band shapes, consistent with minimal differences in respective cysteine side-chain conformations and sulfhydryl interactions. The results imply that neither the Cys 32 nor Cys 35 SH donor is hydrogen bonded directly to Asp 26 in the wild-type protein. Additionally, the thioredoxin main-chain conformation is largely conserved with D26A mutation. Conversely, the mutation perturbs Raman bands diagnostic of tryptophan (Trp 28 and Trp 31) orientations and leads to differences in their pH dependencies, implying local conformational differences near the active site. We conclude that, although the carboxyl side chain of Asp 26 neither interacts directly with active-site cysteines nor is responsible for their abnormally low pKa values, the aspartate side chain may play a role in determining the conformation of the enzyme active site.  相似文献   

8.
To understand the pressure-adaptation mechanism of deep-sea enzymes, we studied the effects of pressure on the enzyme activity and structural stability of dihydrofolate reductase (DHFR) of the deep-sea bacterium Moritella profunda (mpDHFR) in comparison with those of Escherichia coli (ecDHFR). mpDHFR exhibited optimal enzyme activity at 50 MPa whereas ecDHFR was monotonically inactivated by pressure, suggesting inherent pressure-adaptation mechanisms in mpDHFR. The secondary structure of apo-mpDHFR was stable up to 80 °C, as revealed by circular dichroism spectra. The free energy changes due to pressure and urea unfolding of apo-mpDHFR, determined by fluorescence spectroscopy, were smaller than those of ecDHFR, indicating the unstable structure of mpDHFR against pressure and urea despite the three-dimensional crystal structures of both DHFRs being almost the same. The respective volume changes due to pressure and urea unfolding were − 45 and − 53 ml/mol at 25 °C for mpDHFR, which were smaller (less negative) than the corresponding values of − 77 and − 85 ml/mol for ecDHFR. These volume changes can be ascribed to the difference in internal cavity and surface hydration of each DHFR. From these results, we assume that the native structure of mpDHFR is loosely packed and highly hydrated compared with that of ecDHFR in solution.  相似文献   

9.
Acetabularia rhodopsin (AR) is a rhodopsin from the marine plant Acetabularia acetabulum. The opsin-encoding gene from A. acetabulum, ARII, was cloned and found to be novel but homologous to that reported previously. ARII is a light-driven proton pump, as demonstrated by the existence of a photo-induced current through Xenopus oocytes expressing ARII. The photochemical reaction of ARII prepared by cell-free protein synthesis was similar to that of bacteriorhodopsin (BR), except for the lack of light-dark adaptation and the different proton release and uptake sequence. The crystal structure determined at 3.2 Å resolution is the first structure of a eukaryotic member of the microbial rhodopsin family. The structure of ARII is similar to that of BR. From the cytoplasmic side to the extracellular side of the proton transfer pathway in ARII, Asp92, a Schiff base, Asp207, Asp81, Arg78, Glu199, and Ser189 are arranged in positions similar to those of the corresponding residues directly involved in proton transfer by BR. The side-chain carboxyl group of Asp92 appears to interact with the sulfhydryl group of Cys218, which is unique to ARII and corresponds to Leu223 of BR and to Asp217 of Anabaena sensory rhodopsin. The orientation of the Arg78 side chain is opposite to the corresponding Arg82 of BR. The putative absence of water molecules around Glu199 and Arg78 may disrupt the formation of the low-barrier hydrogen bond at Glu199, resulting in the “late proton release”.  相似文献   

10.
N-terminal His-tagged recombinant beta-1,4-galactosyltransferase from Neisseria meningitidis was expressed and purified to homogeneity by column chromatography using Ni-NTA resin. Mutations were introduced to investigate the roles of, Ser68, His69, Glu88, Asp90, and Tyr156, which are components of a highly conserved region in recombinant beta-1,4 galactosyltransferase. Also, the functions of three other cysteine residues, Cys65, Cys139, and Cys205, were investigated using site-directed mutagenesis to determine the location of the disulfide bond and the role of the sulfhydryl groups. Purified mutant galactosyltransferases, His69Phe, Glu88Gln and Asp90Asn completely shut down wild-type galactosyltransferase activity (1-3 %). Also, Ser68Ala showed much lower activity than wild-type galactosyltransferase (19 %). However, only the substitution of Tyr156Phe resulted in a slight reduction in galactosyltransferase activity (90 %). The enzyme was found to remain active when the cysteine residues at positions 139 and 205 were replaced separately with serine. However, enzyme reactivity was found to be markedly reduced when Cys65 was replaced with serine (27 %). These results indicate that conserved amino acids such as Cys65, Ser68, His69, Glu88, and Asp90 may be involved in the binding of substrates or in the catalysis of the galactosyltransferase reaction.  相似文献   

11.
Polyclonal antibodies against dihydrofolate reductase (DHFR) from the human lymphoblastoid cell line WIL-2/M4 were used as probes to compare the antigenic structures in solution of native DHFRs obtained from a broad range of species and their complexes with substrate, cofactor, and folate antagonist inhibitors. All these antibodies could bind to the denatured human DHFR, indicating that they were specific for the primary structure of this enzyme. Denatured chicken liver and L1210 murine leukemic DHFRs competed for all of the antibodies that bound to the human enzyme, although less effectively than the denatured human enzyme, showing the presence of similar epitopes among the vertebrate enzymes. However, both direct binding and competition experiments showed low antibody cross-reactivities with native chicken liver (8%) and murine (10%) DHFRs, suggesting differences in the disposition of similar epitopes in these enzymes. The lactobacillus casei DHFR showed a low amount (less than 2%) of cross-reactivity with the antibodies while the same antibodies did not cross-react with the Escherichia coli enzyme. DHFR from soybean seedlings competed for a large proportion (70%) of the anti-human DHFR antibodies, indicating a close similarity in the antigenic structures of plant and animal DHFRs. Binary complexes of the L. casei, avian, murine, and human DHFRs with dihydrofolate, methotrexate (MTX), trimethoprim (TMP), NADPH, and NADP+ all showed significantly lower antibody binding capacity as compared with the corresponding free enzymes. Further, these ligands inhibited antibody binding to the enzyme to varying degrees.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
The structures of the reaction center variants Pro L209 --> Tyr, Pro L209 --> Phe, and Pro L209 --> Glu from the photosynthetic purple bacterium Rhodobacter sphaeroides have been determined by X-ray crystallography to 2.6-2.8 A resolution. These variants were constructed to interrupt a chain of tightly bound water molecules that was assumed to facilitate proton transfer from the cytoplasm to the secondary quinone Q(B) [Baciou, L., and Michel, H. (1995) Biochemistry 34, 7967-7972]. However, the amino acid exchanges Pro L209 --> Tyr and Pro L209 --> Phe do not interrupt the water chain. Both aromatic side chains are oriented away from this water chain and interact with three surrounding polar side chains (Asp L213, Thr L226, and Glu H173) which are displaced by up to 2.6 A. The conformational changes induced by the bulky aromatic rings of Tyr L209 and Phe L209 lead to unexpected displacements of Q(B) compared to the wild-type protein. In the structure of the Pro L209 --> Tyr variant, Q(B) is shifted by approximately 4 A and is now located at a position similar to that reported for the wild-type reaction center after illumination [Stowell, M. H. B., et al. (1997) Science 276, 812-816]. In the Pro L209 --> Phe variant, the electron density map reveals an intermediate Q(B) position between the binding sites of the wild-type protein in the dark and the Pro L209 --> Tyr protein. In the Pro L209 --> Glu reaction center, the carboxylic side chain of Glu L209 is located within the water chain, and the binding site of Q(B) remains unchanged compared to the wild-type structure.  相似文献   

13.
The functional importance of a conserved region in a novel chitosanase from Bacillus sp. CK4 was investigated. Each of the three carboxylic amino acid residues (Glu-50, Glu-62, and Asp-66) was changed to Asp and Gln or Asn and Glu by site-directed mutagenesis, respectively. The Asp-66-->Asn and Asp-66-->Glu mutation remarkably decreased kinetic parameters such as Vmax and kcat to approximately 1/1,000 those of the wild-type enzyme, indicating that the Asp-66 residue was essential for catalysis. The thermostable chitosanase contains three Cys residues at positions 49, 72, and 211. The Cys-49-->Ser/Tyr and Cys-72-->Ser/Tyr mutant enzymes were as stable to thermal inactivation and denaturating agents as the wild-type enzyme. However, the half-life of the Cys-211-->Ser/Tyr mutant enzyme was less than 10 min at 80 degrees C, while that of the wild-type enzyme was about 90 min. Moreover, the residual activity of Cys-211-->Ser/Tyr enzyme was substantially decreased by 8 M urea; and it lost all catalytic activity in 40% ethanol. These results show that the substitution of Cys with any amino acid residues at position 211 seems to affect the conformational stability of the chitosanase.  相似文献   

14.
Crystal structures of aconitase with isocitrate and nitroisocitrate bound.   总被引:7,自引:0,他引:7  
The crystal structures of mitochondrial aconitase with isocitrate and nitroisocitrate bound have been solved and refined to R factors of 0.179 and 0.161, respectively, for all observed data in the range 8.0-2.1 A. Porcine heart enzyme was used for determining the structure with isocitrate bound. The presence of isocitrate in the crystals was corroborated by M?ssbauer spectroscopy. Bovine heart enzyme was used for determining the structure with the reaction intermediate analogue nitroisocitrate bound. The inhibitor binds to the enzyme in a manner virtually identical to that of isocitrate. Both compounds bind to the unique Fe atom of the [4Fe-4S] cluster via a hydroxyl oxygen and one carboxyl oxygen. A H2O molecule is also bound, making Fe six-coordinate. The unique Fe is pulled away approximately 0.2 A from the corner of the cubane compared to the position it would occupy in a symmetrically ligated [4Fe-4S] cluster. At least 23 residues from all four domains of aconitase contribute to the active site. These residues participate in substrate recognition (Arg447, Arg452, Arg580, Arg644, Gln72, Ser166, Ser643), cluster ligation and interaction (Cys358, Cys421, Cys424, Asn258, Asn446), and hydrogen bonds supporting active site side chains (Ala74, Asp568, Ser571, Thr567). Residues implicated in catalysis are Ser642 and three histidine-carboxylate pairs (Asp100-His101, Asp165-His147, Glu262-His167). The base necessary for proton abstraction from C beta of isocitrate appears to be Ser642; the O gamma atom is proximal to the calculated hydrogen position, while the environment of O gamma suggests stabilization of an alkoxide (an oxyanion hole formed by the amide and side chain of Arg644). The histidine-carboxylate pairs appear to be required for proton transfer reactions involving two oxygens bound to Fe, one derived from solvent (bound H2O) and one derived from substrate hydroxyl. Each oxygen is in contact with a histidine, and both are in contact with the side chain of Asp165, which bridges the two sites on the six-coordinate Fe.  相似文献   

15.
The crystal structures of the four product-complexed single mutants of the catalytic residues of Pseudomonas stutzeri maltotetraose-forming alpha-amylase, E219G, D193N, D193G and D294N, have been determined. Possible roles of the catalytic residues Glu219, Asp193 and Asp294 have been discussed by comparing the structures among the previously determined complexed mutant E219Q and the present mutant enzymes. The results suggested that Asp193 predominantly works as the base catalyst (nucleophile), whose side chain atom lies in close proximity to the C1-atom of Glc4, being involved in the intermediate formation in the hydrolysis reaction. While Asp294 works for tightly binding the substrate to give a twisted and a deformed conformation of the glucose ring at position -1 (Glc4). The hydrogen bond between the side chain atom of Glu219 and the O1-atom of Glc4, that implies the possibility of interaction via hydrogen, consistently present throughout these analyses, supports the generally accepted role of this residue as the acid catalyst (proton donor).  相似文献   

16.
J Heberle  D Oesterhelt    N A Dencher 《The EMBO journal》1993,12(10):3721-3727
Surface bound pH indicators were applied to study the proton transfer reactions in the mutant Asp85-->Glu of bacteriorhodopsin in the native membrane. The amino acid replacement induces a drastic acceleration of the overall rise of the M intermediate. Instead of following this acceleration, proton ejection to the extracellular membrane surface is not only two orders of magnitude slower than M formation, it is also delayed as compared with the wild-type. This demonstrates that Asp85 not only accepts the proton released by the Schiff's base but also regulates very efficiently proton transfer within the proton release chain. Furthermore, Asp85 might be the primary but is not the only proton acceptor/donor group in the release pathway. The Asp85-->Glu substitution also affects the proton reuptake reaction at the cytoplasmic side, although Asp85 is located in the proton release pathway. Proton uptake is slower in the mutant than in the wild-type and occurs during the lifetime of the O intermediate. This demonstrates a feed-back mechanism between Asp85 and the proton uptake pathway in bacteriorhodopsin.  相似文献   

17.
Guan L  Nakae T 《Journal of bacteriology》2001,183(5):1734-1739
The MexABM efflux pump exports structurally diverse xenobiotics, utilizing the proton electrochemical gradient to confer drug resistance on Pseudomonas aeruginosa. The MexB subunit traverses the inner membrane 12 times and has two, two, and one charged residues in putative transmembrane segments 2 (TMS-2), TMS-4, and TMS-10, respectively. All five residues were mutated, and MexB function was evaluated by determining the MICs of antibiotics and fluorescent dye efflux. Replacement of Lys342 with Ala, Arg, or Glu and Glu346 with Ala, Gln, or Asp in TMS-2 did not have a discernible effect. Ala, Asn, or Lys substitution for Asp407 in TMS-4, which is well conserved, led to loss of activity. Moreover, a mutant with Glu in place of Asp407 exhibited only marginal function, suggesting that the length of the side chain at this position is important. The only replacements for Asp408 in TMS-4 or Lys939 in TMS-10 that exhibited significant function were Glu and Arg, respectively, suggesting that the native charge at these positions is required. In addition, double neutral mutants or mutants in which the charged residues Asp407 and Lys939 or Asp408 and Lys939 were interchanged completely lost function. An Asp408-->Glu/Lys939-->Arg mutant retained significant activity, while an Asp407-->Glu/Lys939-->Arg mutant exhibited only marginal function. An Asp407-->Glu/Asp408-->Glu double mutant also lost activity, but significant function was restored by replacing Lys939 with Arg (Asp407-->Glu/Asp408-->Glu/Lys939-->Arg). Taken as a whole, the findings indicate that Asp407, Asp408, and Lys939 are functionally important and raise the possibility that Asp407, Asp408, and Lys939 may form a charge network between TMS-4 and TMS-10 that is important for proton translocation and/or energy coupling.  相似文献   

18.
Clostridium paraputrificum M-21 beta-N-acetylglucosaminidase 3A (Nag3A) is an enzyme classified in family 3 of the glycoside hydrolases. To identify catalytic residues of this enzyme, mutations were introduced into highly conserved Glu and Asp residues. Replacement of Asp175 with Ala abolished the catalytic activity without change in the circular dichroism spectrum, strongly suggesting that this residue is a catalytic residue, a nucleophile/base or a proton donor. Since the K(m) values of mutant enzymes D119N, D229N, D229A and D274N increased 17 to 41 times as compared with that of wild-type enzyme, Asp119, Asp229, and Asp274 appear to be involved in substrate recognition and binding. Taking previous studies into consideration, we presume that Asp303 is the catalytic nucleophile and Asp175 is the proton donor of C. paraputrificum Nag3A.  相似文献   

19.
Montemayor EJ  Hoffman DW 《Biochemistry》2008,47(35):9145-9153
The enzyme spermidine/spermine N (1)-acetyltransferase (SSAT) catalyzes the transfer of acetyl groups from acetylcoenzyme A to spermidine and spermine, as part of a polyamine degradation pathway. This work describes the crystal structure of SSAT in complex with coenzyme A, with and without bound spermine. The complex with spermine provides a direct view of substrate binding by an SSAT and demonstrates structural plasticity near the active site of the enzyme. Associated water molecules bridge several of the intermolecular contacts between spermine and the enzyme and form a "proton wire" between the side chain of Glu92 and the N1 amine of spermine. A single water molecule can also be seen forming hydrogen bonds with the side chains of Glu92, Asp93, and the N4 amine of spermine. Site-directed mutation of Glu92 to glutamine had a detrimental effect on both substrate binding and catalysis and shifted the optimal pH for enzyme activity further into alkaline solution conditions, while mutation of Asp93 to asparagine affected both substrate binding and catalysis without changing the pH dependence of the enzyme. Considered together, the structural and kinetic data suggest that Glu92 functions as a catalytic base to drive an otherwise unfavorable deprotonation step at physiological pH.  相似文献   

20.
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   

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