首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The fatty aldehyde dehydrogenase (Vh-ALDH) isolated from the luminescent bacterium, Vibrio harveyi, differs from other aldehyde dehydrogenases in its high affinity for NADP+. The binding of NADP+ appears to arise from the interaction of the 2′-phosphate of the adenosine moiety of NADP+ with a threonine (T175) in the nucleotide recognition site just after the βB strand as well as with an arginine (R210) that pi stacks over the adenosine moiety. The active site of Vh-ALDH contains the usual suspects of a cysteine (C289), two glutamates (E253 and E377) and an asparagine (N147) involved in the aldehyde dehydrogenase mechanism. However, Vh-ALDH has one polar residue in the active site that distinguishes it from other ALDHs; a histidine (H450) is in close contact with the cysteine nucleophile. As a glutamate has been implicated in promoting the nucleophilicity of the active site cysteine residue in ALDHs, the close contact of a histidine with the cysteine nucleophile in Vh-ALDH raises the possibility of alternate routes to increase the reactivity of the cysteine nucleophile. The effects of mutation of these residues on the different functions catalyzed by Vh-ALDH including acylation, (thio)esterase, reductase and dehydrogenase activities should help define the specific roles of the residues in the active site of ALDHs.  相似文献   

2.
Zhang L  Ahvazi B  Szittner R  Vrielink A  Meighen E 《Biochemistry》2000,39(47):14409-14418
Aldehyde dehydrogenases (ALDHs) catalyze the transfer to NAD(P) of a hydride ion from a thiohemiacetal derivative of the aldehyde coupled with a cysteine residue in the active site. In Vibrio harveyi aldehyde dehydrogenase (Vh-ALDH), a histidine residue (H450) is in proximity (3.8 A) to the cysteine nucleophile (C289) and is thus capable of increasing its reactivity in sharp contrast to other ALDHs in which more distantly located glutamic acid residues are proposed to act as the general base. Mutation of H450 in Vh-ALDH to Gln and Asn resulted in loss of dehydrogenase, (thio)esterase, and acyl-CoA reductase activities; the residual activity of H450Q was higher than that of the H450N mutant in agreement with the capability of Gln but not Asn to partially replace the epsilon-imino group of H450. Coupled with a change in the rate-limiting step, these results indicate that H450 increases the reactivity of C289. Moreover, for the first time, the acylated enzyme intermediate could be directly monitored after reaction with [(3)H]tetradecanoyl-CoA showing that the H450Q mutant was acylated more rapidly than the H450N mutant. Inactivation of the wild-type enzyme with N-ethylmaleimide was much more rapid than the H450Q mutant which in turn was faster than the H450N mutant, demonstrating directly that the nucleophilicity of C289 was affected by H450. As the glutamic acid residue implicated as the general base in promoting cysteine nucleophilicity in other ALDHs is conserved in Vh-ALDH, elucidation of why a histidine residue has evolved to assist in this function in Vh-ALDH will be important to understand the mechanism of ALDHs in general, as well as help delineate the specific roles of the active site glutamic acid residues.  相似文献   

3.
L Zhang  B Ahvazi  R Szittner  A Vrielink  E Meighen 《Biochemistry》1999,38(35):11440-11447
The fatty aldehyde dehydrogenase from the luminescent bacterium, Vibrio harveyi (Vh-ALDH), is unique with respect to its high specificity for NADP(+) over NAD(+). By mutation of a single threonine residue (Thr175) immediately downstream of the beta(B) strand in the Rossmann fold, the nucleotide specificity of Vh-ALDH has been changed from NADP(+) to NAD(+). Replacement of Thr175 by a negatively charged residue (Asp or Glu) resulted in an increase in k(cat)/K(m) for NAD(+) relative to that for NADP(+) of up to 5000-fold due to a decrease for NAD(+) and an increase for NADP(+) in their respective Michaelis constants (K(a)). Differential protection by NAD(+) and NADP(+) against thermal inactivation and comparison of the dissociation constants of NMN, 2'-AMP, 2'5'-ADP, and 5'-AMP for these mutants and the wild-type enzyme clearly support the change in nucleotide specificity. Moreover, replacement of Thr175 with polar residues (N, S, or Q) demonstrated that a more efficient NAD(+)-dependent enzyme T175Q could be created without loss of NADP(+)-dependent activity. Analysis of the three-dimensional structure of Vh-ALDH with bound NADP(+) showed that the hydroxyl group of Thr175 forms a hydrogen bond to the 2'-phosphate of NADP(+). Replacement with glutamic acid or glutamine strengthened interactions with NAD(+) and indicated why threonine would be the preferred polar residue at the nucleotide recognition site in NADP(+)-specific aldehyde dehydrogenases. These results have shown that the size and the structure of the residue at the nucleotide recognition site play the key roles in differentiating between NAD(+) and NADP(+) interactions while the presence of a negative charge is responsible for the decrease in interactions with NADP(+) in Vh-ALDH.  相似文献   

4.
Studies of pH-dependent kinetics implicate two ionizable groups in the dehydrogenase and esterase reactions catalysed by high-Km aldehyde dehydrogenase from rat liver mitochondria. Sensitized photooxidation completely arrests the bifunctional activities of the dehydrogenase. Carboxamidomethylation abolishes the dehydrogenase activity, whereas acetimidination eliminates the esterase activity. These results suggest that histidine (pKa near 6) and cysteine (pKa near 10) are likely the catalytic residues for the dehydrogenase activity, while the esterase activity is functionally related to histidine (pKa near 7) and a residue with the pKa value of 10-11. The two residues, a carboxyl group and an arginine, that discriminate between NAD+ and NADP+ are present at the coenzyme binding site of the mitochondrial high-Km aldehyde dehydrogenase from rat liver.  相似文献   

5.
Relationships within the aldehyde dehydrogenase extended family   总被引:2,自引:0,他引:2       下载免费PDF全文
One hundred-forty-five full-length aldehyde dehydrogenase-related sequences were aligned to determine relationships within the aldehyde dehydrogenase (ALDH) extended family. The alignment reveals only four invariant residues: two glycines, a phenylalanine involved in NAD binding, and a glutamic acid that coordinates the nicotinamide ribose in certain E-NAD binary complex crystal structures, but which may also serve as a general base for the catalytic reaction. The cysteine that provides the catalytic thiol and its closest neighbor in space, an asparagine residue, are conserved in all ALDHs with demonstrated dehydrogenase activity. Sixteen residues are conserved in at least 95% of the sequences; 12 of these cluster into seven sequence motifs conserved in almost all ALDHs. These motifs cluster around the active site of the enzyme. Phylogenetic analysis of these ALDHs indicates at least 13 ALDH families, most of which have previously been identified but not grouped separately by alignment. ALDHs cluster into two main trunks of the phylogenetic tree. The largest, the "Class 3" trunk, contains mostly substrate-specific ALDH families, as well as the class 3 ALDH family itself. The other trunk, the "Class 1/2" trunk, contains mostly variable substrate ALDH families, including the class 1 and 2 ALDH families. Divergence of the substrate-specific ALDHs occurred earlier than the division between ALDHs with broad substrate specificities. A site on the World Wide Web has also been devoted to this alignment project.  相似文献   

6.
The aldehyde dehydrogenases (ALDHs) are a superfamily of multimeric enzymes which catalyse the oxidation of a broad range of aldehydes into their corresponding carboxylic acids with the reduction of their cofactor, NAD or NADP, into NADH or NADPH. At present, the only known structures concern NAD-dependent ALDHs. Three structures are available in the Protein Data Bank: two are tetrameric and the other is a dimer. We solved by molecular replacement the first structure of an NADP-dependent ALDH isolated from Streptococcus mutans, in its apo form and holo form in complex with NADP, at 1.8 and 2.6 A resolution, respectively. Although the protein sequence shares only approximately 30 % identity with the other solved tetrameric ALDHs, the structures are very similar. However, a large local conformational change in the region surrounding the 2' phosphate group of the adenosine moiety is observed when the enzyme binds NADP, in contrast to the NAD-dependent ALDHs.Structure and sequence analyses reveal several properties. A small number of residues seem to determine the oligomeric state. Likewise, the nature (charge and volume) of the residue at position 180 (Thr in ALDH from S. mutans) determines the cofactor specificity in comparison with the structures of NAD-dependent ALDHs. The presence of a hydrogen bond network around the cofactor not only allows it to bind to the enzyme but also directs the side-chains in a correct orientation for the catalytic reaction to take place. Moreover, a specific part of this network appears to be important in substrate binding. Since the enzyme oxidises the same substrate, glyceraldehyde-3-phosphate (G3P), as NAD-dependent phosphorylating glyceraldehyde-3-phosphate dehydrogenases (GAPDH), the active site of GAPDH was compared with that of the S. mutans ALDH. It was found that Arg103, Arg283 and Asp440 might be key residues for substrate binding.  相似文献   

7.
The enzyme, 10-formyltetrahydrofolate dehydrogenase (FDH), converts 10-formyltetrahydrofolate (10-formyl-THF) to tetrahydrofolate in an NADP(+)-dependent dehydrogenase reaction or an NADP(+)-independent hydrolase reaction. The hydrolase reaction occurs in a 310-amino acid long amino-terminal domain of FDH (N(t)-FDH), whereas the dehydrogenase reaction requires the full-length enzyme. The amino-terminal domain of FDH shares some sequence identity with several other enzymes utilizing 10-formyl-THF as a substrate. These enzymes have two strictly conserved residues, aspartate and histidine, in the putative catalytic center. We have shown recently that the conserved aspartate is involved in FDH catalysis. In the present work we studied the role of the conserved histidine, His(106), in FDH function. Site-directed mutagenesis experiments showed that replacement of the histidine with alanine, asparagine, aspartate, glutamate, glutamine, or arginine in N(t)-FDH resulted in expression of insoluble proteins. Replacement of the histidine with another positively charged residue, lysine, produced a soluble mutant with no hydrolase activity. The insoluble mutants refolded from inclusion bodies adopted a conformation inherent to the wild-type N(t)-FDH, but they did not exhibit any hydrolase activity. Substitution of alanine for three non-conserved histidines located close to the conserved one did not reveal any significant changes in the hydrolase activity of N(t)-FDH. Expressed full-length FDH with the substitution of lysine for the His(106) completely lost both the hydrolase and dehydrogenase activities. Thus, our study showed that His(106), besides being an important structural residue, is also directly involved in both the hydrolase and dehydrogenase mechanisms of FDH. Modeling of the putative hydrolase catalytic center/folate-binding site suggested that the catalytic residues, aspartate and histidine, are unlikely to be adjacent to the catalytic cysteine in the aldehyde dehydrogenase catalytic center. We hypothesize that 10-formyl-THF dehydrogenase reaction is not an independent reaction but is a combination of hydrolase and aldehyde dehydrogenase reactions.  相似文献   

8.
The NAD(+)-dependent non-phosphorylating glyceraldehyde-3-phosphate dehydrogenase (GAPN) from the hyperthermophilic archaeum Thermoproteus tenax represents an archaeal member of the diverse superfamily of aldehyde dehydrogenases (ALDHs). GAPN catalyzes the irreversible oxidation of d-glyceraldehyde 3-phosphate to 3-phosphoglycerate. In this study, we present the crystal structure of GAPN in complex with its natural inhibitor NADP(+) determined by multiple anomalous diffraction methods. The structure was refined to a resolution of 2.4 A with an R-factor of 0.21. The overall fold of GAPN is similar to the structures of ALDHs described previously, consisting of three domains: a nucleotide-binding domain, a catalytic domain, and an oligomerization domain. Local differences in the active site are responsible for substrate specificity. The inhibitor NADP(+) binds at an equivalent site to the cosubstrate-binding site of other ALDHs and blocks the enzyme in its inactive state, possibly preventing the transition to the active conformation. Structural comparison between GAPN from the hyperthermophilic T. tenax and homologs of mesophilic organisms establishes several characteristics of thermostabilization. These include protection against heat-induced covalent modifications by reducing and stabilizing labile residues, a decrease in number and volume of empty cavities, an increase in beta-strand content, and a strengthening of subunit contacts by ionic and hydrophobic interactions.  相似文献   

9.
In sorbitol dehydrogenase only one cysteine residue, Cys-43, is reactive in both anionic buffer (phosphate) and zinc-liganding buffer (imidazole) upon carboxymethylation. This is in contrast to the situation in the structurally related liver alcohol dehydrogenase, with either of two alternative Cys residues being reactive, and is compatible with differences in zinc-binding and active site relationships between these two metalloenzymes. Unrelated aldehyde dehydrogenase, upon carboxamidomethylation, shows a third pattern, now less well defined but confirming the presence of a thiol function of Cys-302 close to the active site.  相似文献   

10.
Concentrated urine formation in the kidney is accompanied by conditions that favor the accumulation of reactive oxygen species (ROS). Under hyperosmotic conditions, medulla cells accumulate glycine betaine, which is an osmolyte synthesized by betaine aldehyde dehydrogenase (BADH, EC 1.2.1.8). All BADHs identified to date have a highly reactive cysteine residue at the active site, and this cysteine is susceptible to oxidation by hydrogen peroxide. Porcine kidney BADH incubated with H(2)O(2) (0-500 μM) lost 25% of its activity. However, pkBADH inactivation by hydrogen peroxide was limited, even after 120 min of incubation. The presence of coenzyme NAD(+) (10-50 μM) increased the extent of inactivation (60%) at 120 min of reaction, but the ligands betaine aldehyde (50 and 500 μM) and glycine betaine (100 mM) did not change the rate or extent of inactivation as compared to the reaction without ligand. 2-Mercaptoethanol and dithiothreitol, but not reduced glutathione, were able to restore enzyme activity. Mass spectrometry analysis of hydrogen peroxide inactivated BADH revealed oxidation of M278, M243, M241 and H335 in the absence and oxidation of M94, M327 and M278 in the presence of NAD(+). Molecular modeling of BADH revealed that the oxidized methionine and histidine residues are near the NAD(+) binding site. In the presence of the coenzyme, these oxidized residues are proximal to the betaine aldehyde binding site. None of the oxidized amino acid residues participates directly in catalysis. We suggest that pkBADH inactivation by hydrogen peroxide occurs via disulfide bond formation between vicinal catalytic cysteines (C288 and C289).  相似文献   

11.
Influences on coenzyme preference are explored. Lysine 137 (192 in class 1/2 ALDH) lies close to the adenine ribose, directly interacting with the adenine ribose in NAD-specific ALDHs and the 2'-phosphate of NADP in NADP-specific ALDHs. Lys-137 in class 3 ALDH interacts with the adenine ribose indirectly through an intervening water molecule. However, this residue is present in all ALDHs and, as a result, is unlikely to directly influence coenzyme specificity. Glutamate 140 (195) coordinates the 2'- and 3'-hydroxyls of the adenine ribose of NAD in the class 3 tertiary structure. Thus, it appeared that this residue would influence coenzyme specificity. Mutation to aspartate, asparagine, glutamine or threonine shifts the coenzyme specificity towards NADP, but did not completely change the specificity. Still, the mutants show the 2'-phosphate of NADP is repelled by Glu-140 (195). Although Glu-140 (195) has a major influence on coenzyme specificity, it is not the only influence since class 3 ALDHs, can use both coenzymes, and class 2 ALDHs, which are NAD-specific, have a glutamate at this position. One explanation may be that the larger space between Lys-137 (192) and the adenine ribose hydroxyls in the class 3 ALDH:NAD binary structure may provide space to accommodate the 2'-phosphate of NADP. Also, a structural shift upon binding NADP may also occur in class 3 ALDHs to help accommodate the 2'-phosphate of NADP.  相似文献   

12.
Influences on coenzyme preference are explored. Lysine 137 (192 in class 1/2 ALDH) lies close to the adenine ribose, directly interacting with the adenine ribose in NAD-specific ALDHs and the 2′-phosphate of NADP in NADP-specific ALDHs. Lys-137 in class 3 ALDH interacts with the adenine ribose indirectly through an intervening water molecule. However, this residue is present in all ALDHs and, as a result, is unlikely to directly influence coenzyme specificity. Glutamate 140 (195) coordinates the 2′- and 3′-hydroxyls of the adenine ribose of NAD in the class 3 tertiary structure. Thus, it appeared that this residue would influence coenzyme specificity. Mutation to aspartate, asparagine, glutamine or threonine shifts the coenzyme specificity towards NADP, but did not completely change the specificity. Still, the mutants show the 2′-phosphate of NADP is repelled by Glu-140 (195). Although Glu-140 (195) has a major influence on coenzyme specificity, it is not the only influence since class 3 ALDHs, can use both coenzymes, and class 2 ALDHs, which are NAD-specific, have a glutamate at this position. One explanation may be that the larger space between Lys-137 (192) and the adenine ribose hydroxyls in the class 3 ALDH:NAD binary structure may provide space to accommodate the 2′-phosphate of NADP. Also, a structural shift upon binding NADP may also occur in class 3 ALDHs to help accommodate the 2′-phosphate of NADP.  相似文献   

13.
The mitochondrial isozyme of horse liver aldehyde dehydrogenase was labeled with brominated [5-(3-acetylpyridinio)pentyl]diphosphoadenosine. Specific labeling of a coenzyme binding region was proven by an enzymatic activity of the isozyme with the nonbrominated coenzyme derivative, optical properties of the complex, stoichiometry of incorporation, and protection against inactivation. A cysteine residue was selectively modified by the brominated coenzyme analogue and was identified in a 35-residue tryptic peptide. This cysteine residue corresponds to Cys-302 of the cytoplasmic isozyme and has earlier been implicated in disulfiram binding, confirming a position close to the active site. In contrast, the butyl homologue of the coenzyme analogue labels another residue of the mitochondrial isozyme. Thus, in the same isozyme, two residues are selectively reactive. They are concluded to be close together in the tertiary structure and to be close enough to the coenzyme binding site to be differentially labeled by coenzyme analogues differing only by a single methylene group.  相似文献   

14.
Lipid peroxidation is one of the consequences of environmental stress in plants and leads to the accumulation of highly toxic, reactive aldehydes. One of the processes to detoxify these aldehydes is their oxidation into carboxylic acids catalyzed by NAD(P)+-dependent ALDHs (aldehyde dehydrogenases). We investigated kinetic parameters of two Arabidopsis thaliana family 3 ALDHs, the cytosolic ALDH3H1 and the chloroplastic isoform ALDH3I1. Both enzymes had similar substrate specificity and oxidized saturated aliphatic aldehydes. Catalytic efficiencies improved with the increase of carbon chain length. Both enzymes were also able to oxidize α,β-unsaturated aldehydes, but not aromatic aldehydes. Activity of ALDH3H1 was NAD+-dependent, whereas ALDH3I1 was able to use NAD+ and NADP+. An unusual isoleucine residue within the coenzyme-binding cleft was responsible for the NAD+-dependence of ALDH3H1. Engineering the coenzyme-binding environment of ALDH3I1 elucidated the influence of the surrounding amino acids. Enzyme activities of both ALDHs were redox-sensitive. Inhibition was correlated with oxidation of both catalytic and non-catalytic cysteine residues in addition to homodimer formation. Dimerization and inactivation could be reversed by reducing agents. Mutant analysis showed that cysteine residues mediating homodimerization are located in the N-terminal region. Modelling of the protein structures revealed that the redox-sensitive cysteine residues are located at the surfaces of the subunits.  相似文献   

15.
The C-terminal domain (C(t)-FDH) of 10-formyltetrahydrofolate dehydrogenase (FDH, ALDH1L1) is an NADP(+)-dependent oxidoreductase and a structural and functional homolog of aldehyde dehydrogenases. Here we report the crystal structures of several C(t)-FDH mutants in which two essential catalytic residues adjacent to the nicotinamide ring of bound NADP(+), Cys-707 and Glu-673, were replaced separately or simultaneously. The replacement of the glutamate with an alanine causes irreversible binding of the coenzyme without any noticeable conformational changes in the vicinity of the nicotinamide ring. Additional replacement of cysteine 707 with an alanine (E673A/C707A double mutant) did not affect this irreversible binding indicating that the lack of the glutamate is solely responsible for the enhanced interaction between the enzyme and the coenzyme. The substitution of the cysteine with an alanine did not affect binding of NADP(+) but resulted in the enzyme lacking the ability to differentiate between the oxidized and reduced coenzyme: unlike the wild-type C(t)-FDH/NADPH complex, in the C707A mutant the position of NADPH is identical to the position of NADP(+) with the nicotinamide ring well ordered within the catalytic center. Thus, whereas the glutamate restricts the affinity for the coenzyme, the cysteine is the sensor of the coenzyme redox state. These conclusions were confirmed by coenzyme binding experiments. Our study further suggests that the binding of the coenzyme is additionally controlled by a long-range communication between the catalytic center and the coenzyme-binding domain and points toward an α-helix involved in the adenine moiety binding as a participant of this communication.  相似文献   

16.
Among pyridine-nucleotide-dependent oxidoreductases, the class 3 family of aldehyde dehydrogenases (ALDHs) is unusual in its ability to function with either NAD or NADP. This is all the more surprising because an acidic residue, Glu140, coordinates the adenine ribose 2' hydroxyl. In many NAD-dependent dehydrogenases a similarly placed carboxylate is thought to be responsible for exclusion of NADP. The corresponding residue in most (approximately 71%) sequences in the ALDH extended family is also Glu, and most of these are NAD-specific enzymes. Site-directed mutagenesis was performed on this residue in rat class 3 ALDH. Our results indicate that this residue contributes to tighter binding of NAD in the native enzyme, but suggest that additional factors must contribute to the ability to utilize NADP. Mutagenesis of an adjacent basic residue (Lys137) indicates that it is even more essential for binding both coenzymes, consistent with its conservation in nearly all ALDHs (> 98%).  相似文献   

17.
Alignment of all known, diverse members of the aldehyde dehydrogenase (ALDH) extended family revealed only two strictly conserved, nonglycine residues, a glutamate and a phenylalanine residue. Both occur in one of the highly conserved 'motif' segments and both occupy strategic locations in the tertiary structure at the bottom of the catalytic funnel. In class 3 ALDH, these are Glu333 and Phe335. In addition, Asp247, which is not highly conserved but is characteristic of class 3 ALDHs, hydrogen bonds the main chain between Glu333 and Phe335. These three residues were mutated conservatively. Michaelis constants determined for both NAD/propanal and NADP/benzaldehyde substrate pairs show all three residues to be crucial to effective catalysis, and suggest that the hydrogen bond to Asp247 is a key element in maintaining precise geometry of key elements at the active site.  相似文献   

18.
Dihydrodiol dehydrogenase from pig liver was inactivated by diethylpyrocarbonate (DEP) and by rose bengal-sensitized photooxidation. The DEP inactivation was reversed by hydroxylamine and the absorption spectrum of the inactivated enzyme indicated that both histidine and tyrosine residues were carbethoxylated. The rates of inactivation by DEP and by photooxidation were dependent on pH, showing the involvement of a group with a pKa of 6.4. The kinetics of inactivation and spectrophotometric quantification of the modified residues suggested that complete inactivation was caused by modification of one histidine residue per active site. The inactivation by the two modifications was partially prevented by either NADP(H) or the combination of NADP+ and substrate, and completely prevented in the presence of both NADP+ and a competitive inhibitor which binds to the enzyme-NADP+ binary complex. The DEP-modified enzyme caused the same blue shift and enhancement of NADPH fluorescence as did the native enzyme, suggesting that the modified histidine is not in the coenzyme-binding site of the enzyme. The results suggest the presence of essential histidine residues in the catalytic region of the active site of pig liver dihydrodiol dehydrogenase.  相似文献   

19.
Human liver cytosolic and mitochondrial isozymes of aldehyde dehydrogenase share 70% sequence identity. However, the first 21 residues are not conserved between the human isozymes (15% identity). The three-dimensional structures of the beef mitochondrial and sheep cytosolic forms have virtually identical three-dimensional structures. Here, we solved the structure of the human mitochondrial enzyme and found it to be identical to the beef enzyme. The first 21 residues are found on the surface of the enzyme and make no contact with other subunits in the tetramer. A pair of chimeric enzymes between the human isozymes was made. Each chimera had the first 21 residues from one isozyme and the remaining 479 from the other. When the first 21 residues were from the mitochondrial isozyme, an enzyme with cytosolic-like properties was produced. The other was expressed but was insoluble. It was possible to restore solubility and activity to the chimera that had the first 21 cytosolic residues fused to the mitochondrial ones by making point mutations to residues at the N-terminal end. When residue 19 was changed from tyrosine to a cysteine, the residue found in the mitochondrial form, an active enzyme could be made though the Km for NAD+ was 35 times higher than the native mitochondrial isozyme and the specific activity was reduced by 75%. This residue interacts with residue 203, a nonconserved, nonactive site residue. A mutation of residue 18, which also interacts with 203, restored solubility, but not activity. Mutation to residue 15, which interacts with 104, also restored solubility but not activity. It appears that to have a soluble or active enzyme a favorable interaction must occur between a residue in a surface loop and a residue elsewhere in the molecule even though neither make contact with the active site region of the enzyme.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号