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 46 毫秒
1.
Carere J  Baker P  Seah SY 《Biochemistry》2011,50(39):8407-8416
BphI-BphJ, an aldolase-dehydrogenase complex from the polychlorinated biphenyls (PCBs) degradation pathway, cleaves 4-hydroxy-2-oxoacids to pyruvate and an aldehyde. The enzyme complex was shown to exhibit substrate channeling, whereby linear aldehydes of up to 6 carbons long and branched isobutyraldehyde were directly channeled from the aldolase to the dehydrogenase with greater than 80% efficiency. BphI variants G322F, G322L, and G323F were created and were found to block aldehyde channeling. The dehydrogenase cofactor NADH was able to activate the catalytic activity of the aldol cleavage reaction in these variants, suggesting that activation of BphI by BphJ cofactors is not solely due to faster aldehyde release. A G323L variant was able to channel acetaldehyde but not the larger propionaldehyde while the G323A variant was able to channel butyraldehyde but not its isomer isobutyraldehyde, confirming that the restricted channeling of aldehydes in these glycine variants are due to steric blockage of the channel. Substitution of His-20 and Tyr-290 in BphI led to significant reductions in aldehyde channeling efficiencies. A mechanism of substrate channeling involving these two gating residues is proposed.  相似文献   

2.
Baker P  Hillis C  Carere J  Seah SY 《Biochemistry》2012,51(9):1942-1952
Bacterial aldolase-dehydrogenase complexes catalyze the last steps in the meta cleavage pathway of aromatic hydrocarbon degradation. The aldolase (TTHB246) and dehydrogenase (TTHB247) from Thermus thermophilus were separately expressed and purified from recombinant Escherichia coli. The aldolase forms a dimer, while the dehydrogenase is a monomer; these enzymes can form a stable tetrameric complex in vitro, consisting of two aldolase and two dehydrogenase subunits. Upon complex formation, the K(m) value of 4-hydroxy-2-oxopentanoate, the substrate of TTHB246, is decreased 4-fold while the K(m) of acetaldehyde, the substrate of TTHB247, is increased 3-fold. The k(cat) values of each enzyme were reduced by ~2-fold when they were in a complex. The half-life of TTHB247 at 50 °C increased by ~4-fold when it was in a complex with TTHB246. The acetaldehyde product from TTHB246 could be efficiently channelled directly to TTHB247, but the channeling efficiency for the larger propionaldehyde was ~40% lower. A single A324G substitution in TTHB246 increased the channeling efficiency of propionaldehyde to a value comparable to that of acetaldehyde. Stable and catalytically competent chimeric complexes could be formed between the T. thermophilus enzymes and the orthologous aldolase (BphI) and dehydrogenase (BphJ) from the biphenyl degradation pathway of Burkholderia xenovorans LB400. However, channeling efficiencies for acetaldehyde in these chimeric complexes were ~10%. Structural and sequence analysis suggests that interacting residues in the interface of the aldolase-dehydrogenase complex are highly conserved among homologues, but coevolution of partner enzymes is required to fine-tune this interaction to allow for efficient substrate channeling.  相似文献   

3.
Baker P  Carere J  Seah SY 《Biochemistry》2011,50(17):3559-3569
BphI, a pyruvate-specific class II aldolase found in the polychlorinated biphenyls (PCBs) degradation pathway, catalyzes the reversible C-C bond cleavage of (4S)-hydroxy-2-oxoacids to form pyruvate and an aldehyde. Mutations were introduced into bphI to probe the contribution of active site residues to substrate recognition and catalysis. In contrast to the wild-type enzyme that has similar specificities for acetaldehyde and propionaldehyde, the L87A variant exhibited a 40-fold preference for propionaldehyde over acetaldehyde. The specificity constant of the L89A variant in the aldol addition reaction using pentaldehyde is increased ~50-fold, making it more catalytically efficient for pentaldehyde utilization compared to the wild-type utilization of the natural substrate, acetaldehyde. Replacement of Tyr-290 with phenylalanine or serine resulted in a loss of stereochemical control as the variants were able to utilize substrates with both R and S configurations at C4 with similar kinetic parameters. Aldol cleavage and pyruvate α-proton exchange activity were undetectable in the R16A variant, supporting the role of Arg-16 in stabilizing a pyruvate enolate intermediate. The pH dependence of the enzyme is consistent with a single deprotonation by a catalytic base with pK(a) values of approximately 7. In H20A and H20S variants, pH profiles show the dependence of enzyme activity on hydroxide concentration. On the basis of these results, a catalytic mechanism is proposed.  相似文献   

4.
A study was made of the effect of chronic administration of the hypolipidemic drug clofibrate on the activity and intracellular localization of rat liver aldehyde dehydrogenase. The enzyme was assayed using several aliphatic and aromatic aldehydes. Clofibrate treatment caused a 1.5 to 2.3-fold increase in the liver specific aldehyde dehydrogenase activity. The induced enzyme has a high Km for acetaldehyde and was found to be located in peroxisomes and microsomes. Clofibrate did not alter the enzyme activity in the cytoplasmic fraction. The total peroxisomal aldehyde dehydrogenase activity increased 3 to 4-fold under the action of clofibrate. Disruption of the purified peroxisomes by the hypotonic treatment or in the alkaline conditions resulted in the release of catalase from the broken organelles, while aldehyde dehydrogenase as well as nucleoid-bound urate oxidase and the peroxisomal membrane marker NADH:cytochrome c reductase remained in the peroxisomal 'ghosts'. At the same time, treatment by Triton X-100 led to solubilization of the membrane-bound NADH:cytochrome c reductase and aldehyde dehydrogenase from intact peroxisomes and their 'ghosts'. These results indicate that aldehyde dehydrogenase is located in the peroxisomal membrane. The peroxisomal aldehyde dehydrogenase is active with different aliphatic and aromatic aldehydes, except for formaldehyde and glyceraldehyde. The enzyme Km values lie in the millimolar range for acetaldehyde, propionaldehyde, benzaldehyde and phenylacetaldehyde and in the micromolar range for nonanal. Both NAD and NADP serve as coenzymes for the enzyme. Aldehyde dehydrogenase was inhibited by disulfiram, N-ethylmaleimide and 5,5'-dithiobis(2-nitrobenzoic)acid. According to its basic kinetic properties peroxisomal aldehyde dehydrogenase seems to be similar to a clofibrate-induced microsomal enzyme. The functional role of both enzymes in the liver cells is discussed.  相似文献   

5.
Two isozymes (E1 and E2) of human aldehyde dehydrogenase (EC 1.2.1.3) were purified to homogeneity 13 years ago and a third isozyme (E3) with a low Km for gamma-aminobutyraldehyde only recently. Comparison with a variety of substrates demonstrates that substrate specificity of all three isozymes is broad and similar. With straight chain aliphatic aldehydes (C1-C6) the Km values of the E3 isozyme are identical with those of the E1 isozyme. All isozymes dehydrogenate naturally occurring aldehydes, 5-imidazoleacetaldehyde (histamine metabolite) and acrolein (product of beta-elimination of oxidized polyamines) with similar catalytic efficiency. Differences between the isozymes are in the Km values for aminoaldehydes. Although all isozymes can dehydrogenate gamma-aminobutyraldehyde, the Km value of the E3 isozyme is much lower: the same appears to apply to aldehyde metabolites of cadaverine, agmatine, spermidine, and spermine for which Km values range between 2-18 microM and kcat values between 0.8-1.9 mumol/min/mg. Thus, the E3 isozyme has properties which make it suitable for the metabolism of aminoaldehydes. The physiological role of E1 and E2 isozymes could be in dehydrogenation of aldehyde metabolites of monoamines such as 3,4-dihydroxyphenylacetaldehyde or 5-hydroxyindoleacetaldehyde; the catalytic efficiency with these substrates is better with E1 and E2 isozymes than with E3 isozyme. Isoelectric focusing of liver homogenates followed by development with various physiological substrates together with substrate specificity data suggest that aldehyde dehydrogenase (EC 1.2.1.3) is the only enzyme in the human liver capable of catalyzing dehydrogenation of aldehydes arising via monoamine, diamine, and plasma amine oxidases. Although the enzyme is generally considered to function in detoxication, our data suggest an additional function in metabolism of biogenic amines.  相似文献   

6.
Some physical and chemical properties of the monomeric NADP+-dependent aldehyde reductase (previously called TPN-L-hexonate dehydrogenase or D-glucuronate reductase) from pig kidney have been examined. The amino acid composition has been determined. Four of the five thiol groups react with p-mercuribenzoate at pH 7, with no resulting loss of catalytic activity. High concentrations of p-mercuribenzoate cause complete enzyme inhibition, which can be partly reversed by addition of aldehyde reductase is low (9%, estimated from the ellipticity at 208 nm), and 70 to 80% of the tyrosine and tryptophan residues aare buried within the molecule. One molecule of NADPH binds to the enzyme (Kp equal 25 muM), causing a blue shift and enhancement of the coenzyme fluorescence, and suggesting that the environment of the active site is hydrophobic. In the reduction of D-glyceraldehyde, catalyzed by aldehyde reductase, the pro-4R "A" hydrogen of NADPH attacks the re face of the carbonyl group. This stereospecificity is the same as in the reductions of D-glyceraldehyde and acetaldehyde effected by rabbit muscle dehydrogenase and liver alcohol dehydrogenase, respectively.  相似文献   

7.
Experiments were carried out to study the effect of two commonly used glutathione-depleting agents, diethylmaleate and phorone, on the oxidation of acetaldehyde and the activity of aldehyde dehydrogenase. The oxidation of acetaldehyde by intact hepatocytes was inhibited when the cells were incubated with diethylmaleate. Washing and resuspending the cells in diethylmaleate-free medium afforded protection against the inhibition of acetaldehyde oxidation. The oxidation of acetaldehyde by isolated rat liver mitochondria as well as by disrupted mitochondria in the presence of excess NAD+ was inhibited by diethylmaleate or phorone, indicating inhibition of the low-Km aldehyde dehydrogenase. In addition, diethylmaleate inhibited oxidation of acetaldehyde by the high-Km cytosolic aldehyde dehydrogenase. Significant accumulation of acetaldehyde occurred when ethanol was oxidized by hepatocytes in the presence, but not in the absence, of diethylmaleate. Thus, diethylmaleate blocks the oxidation of added or metabolically generated acetaldehyde, analogous to results with other inhibitors of the low-Km aldehyde dehydrogenase such as cyanamide. These results suggest that caution should be used in interpreting the effects of diethylmaleate or phorone on metabolic reactions, especially those involving metabolism of aldehydes such as formaldehyde, because, in addition to depleting glutathione, these agents inhibit the low-Km aldehyde dehydrogenase.  相似文献   

8.
Ethanol or acetaldehyde orally administered (15% and 2% respectively in drinking water) to male Wistar rats for three months induced alterations in the main liver enzymes responsible for ethanol metabolism, aspartate and alanine aminotransferases and NAD glutamate dehydrogenase. Ethanol produced a significant decrease in the activity of soluble alcohol dehydrogenase, while acetaldehyde induced alterations both in soluble and mitochondrial aldehyde dehydrogenases: soluble activity was significantly higher than in the control and ethanol-treated groups, and mitochondrial activity was significantly diminished. Both soluble aspartate and alanine aminotransferases showed pronounced increases by the chronic effect of acetaldehyde, while mitochondrial activities were practically unchanged by the effect of ethanol or acetaldehyde. Mitochondrial NAD glutamate dehydrogenase showed a rise in its activity both by the effect of chronic ethanol and acetaldehyde consumption. The level of metabolites assayed in liver extracts showed marked differences between ethanol and acetaldehyde treatment which indicates that ethanol produced a remarkable increase in glutamate, aspartate and free ammonia together with marked decrease in pyruvate and 2-oxoglutarate concentrations. Acetaldehyde consumption induced a significant decrease in 2-oxoglutarate and pyruvate concentrations. These observations suggest that ethanol has an important effect on the urea cycle enzymes, while the effect of acetaldehyde contributes to the impairment of the citric acid cycle.  相似文献   

9.

2-Deoxy-d-ribose-5-phosphate aldolase (DERA) is a class I aldolase that offers access to several building blocks for organic synthesis. It catalyzes the stereoselective C–C bond formation between acetaldehyde and numerous other aldehydes. However, the practical application of DERA as a biocatalyst is limited by its poor tolerance towards industrially relevant concentrations of aldehydes, in particular acetaldehyde. Therefore, the development of proper experimental conditions, including protein engineering and/or immobilization on appropriate supports, is required. The present review is aimed to provide a brief overview of DERA, its history, and progress made in understanding the functioning of the enzyme. Furthermore, the current understanding regarding aldehyde resistance of DERA and the various optimizations carried out to modify this property are discussed.

  相似文献   

10.
Ho KK  Weiner H 《Journal of bacteriology》2005,187(3):1067-1073
An aldehyde dehydrogenase was detected in crude cell extracts of Escherichia coli DH5alpha. Growth studies indicated that the aldehyde dehydrogenase activity was growth phase dependent and increased in cells grown with ethanol. The N-terminal amino acid sequence of the purified enzyme identified the latter as an aldehyde dehydrogenase encoded by aldB, which was thought to play a role in the removal of aldehydes and alcohols in cells that were under stress. The purified enzyme showed an estimated molecular mass of 220 +/- 8 kDa, consisting of four identical subunits, and preferred to use NADP and acetaldehyde. MgCl2 increased the activity of the NADP-dependent enzyme with various substrates. A comparison of the effect of Mg2+ ions on the bacterial enzyme with the effect of Mg2+ ions on human liver mitochondrial aldehyde dehydrogenase revealed that the bacterial enzyme shared kinetic properties with the mammalian enzyme. An R197E mutant of the bacterial enzyme appeared to retain very little NADP-dependent activity on acetaldehyde.  相似文献   

11.
Summary In the final step of the pathway producing ethanol in anoxic crucian carp (Carassius carassius L.), acetaldehyde is reduced to ethanol by alcohol dehydrogenase. The presence of aldehyde dehydrogenase in the tissues responsible for ethanol production could cause an undesired oxidation of acetaldehyde to acetate coupled with a reduction of NAD+ to NADH. Moreover, acetaldehyde could competitively inhibit the oxidation of reactive biogenic aldehydes. In the present study, the distribution of aldehyde dehydrogenase (measured with a biogenic aldehyde) and alcohol dehydrogenase (measured with acetaldehyde) were studied in organs of crucian carp, common carp (Cyprinus carpio L.), rainbow trout (Salmo gairdneri Richardson), and Norwegian rat (Rattus norvegicus Berkenhout). The results showed that alcohol dehydrogenase and aldehyde dehydrogenase activities were almost completely spatially separated in the crucian carp. These enzymes occurred together in the other three vertebrates. In the crucian carp, alcohol dehydrogenase was only found in red and white skeletal muscle, while these tissues contained exceptionally low aldehyde dehydrogenase activities. Moreover, the low aldehyde dehydrogenase activity found in crucian carp red muscle was about 1000 times less sensitive to inhibition by acetaldehyde than that found in other tissues and other species. The results are interpreted as demonstrating adaptations to avoid a depletion of ethanol production, and possibly inhibition of biogenic aldehyde metabolism.Abbreviations ADH alcohol dehydrogenase - ALDH aldehyde dehydrogenase - DOPAL 3,4-dihydroxyphenylacetaldehyde - MAO monoamine oxidase - PCA perchloric acid  相似文献   

12.
Hyphomicrobium strain WC, Pseudomonas strain TP-1, and Pseudomonas strain W1 are capable of growth on methanol as the sole source of carbon and energy. Methanol-grown cells of each organism contain a primary alcohol dehydrogenase that has been purified to homogeneity. Each enzyme has a molecular weight of 120,000 and shows an in vitro requirement for phenazine methosulfate and ammonium ions for enzymatic activity. Normal aliphatic alcohols are oxidized rapidly by each enzyme. The presence of a methyl group on the carbon atom adjacent to the primary alcohol group lowers the enzymatic activity. This effect is reduced as the methyl substituent is moved further away from the hydroxyl group. The effect of other substituents on enzymatic activity is reported. Methanol, formaldehyde, and to a limited extent acetaldehyde are oxidized by the primary alcohol dehydrogenases. Higher aldehydes are not oxidized. A possible explanation for this specificity, with regard to aldehydes, is presented in terms of degree of hydration of the aldehyde.  相似文献   

13.
Brownian dynamics (BD) simulations test for channeling of the substrate, glyceraldehyde 3-phosphate (GAP), as it passes between the enzymes fructose-1,6-bisphosphate aldolase (aldolase) and glyceraldehyde 3-phosphate dehydrogenase (GAPDH). First, BD simulations determined the favorable complexes between aldolase and GAPDH; two adjacent subunits of GAPDH form salt bridges with two subunits of aldolase. These intermolecular contacts provide a strong electrostatic interaction between the enzymes. Second, BD simulates GAP moving out of the active site of the A or D aldolase subunit and entering any of the four active sites of GAPDH. The efficiency of transfer is determined as the relative number of BD trajectories that reached any active site of GAPDH. The distribution functions of the transfer time were calculated based on the duration of successful trajectories. BD simulations of the GAP binding from solution to aldolase/GAPDH complex were compared to the channeling simulations. The efficiency of transfer of GAP within an aldolase/GAPDH complex was 2 to 3% compared to 1.3% when GAP was binding to GAPDH from solution. There is a preference for GAP channeling between aldolase and GAPDH when compared to binding from solution. However, this preference is not large enough to be considered as a theoretical proof of channeling between these proteins.  相似文献   

14.
A novel lyase, combining hydratase and aldolase activity, that converts citral into methylheptenone and acetaldehyde, was purified from spores of Penicillium digitatum. Remarkably, citral lyase activity was induced 118-fold by incubating nongerminating spores with the substrate, citral. This cofactor independent hydratase/aldolase, was purified and found to be a monomeric enzyme of 31 kDa. Citral lyase has a Km of 0.058 mm and a Vmax of 52.6 U.mg-1. Enzyme activity was optimal at 20 degrees C and pH 7.6. The enzyme has a strong preference for the trans isomer of citral (geranial). Citral lyase also converts other alpha,beta-unsaturated aldehydes (farnesal, methyl-crotonaldehyde, decenal and cinnemaldehyde).  相似文献   

15.
Aldehyde dehydrogenase was purified 187-fold from cytosol of rat testis by chromatographic methods and gel filtration with a yield of about 50%. The enzyme exhibits absolute requirement for exogenous sulfhydryl compounds and strong dependence on temperature. Addition of 0.4mM Ca2 or Mg2 ions results in 50% inhibition. Optimally active at pH 8.5 and 50 degrees C, aldehyde dehydrogenase displays broad substrate specificity; saturation curves with acetaldehyde and propionaldehyde are non-hyperbolic, with Hill coefficients comprised between 0.8 and 0.7. Strong substrate inhibition can be observed with both aromatic and long-chain alyphatic aldehydes. According to mathematical models, Km decreases from 246 microM for acetaldehyde to 4 microM for capronaldehyde and Ki decreases from about 4mM for butyraldehyde to 0.2 mM for capronaldehyde.  相似文献   

16.
The enzymes catalyzing the NAD-dependent oxidation of malonic dialdehyde (MDA) were isolated from rat liver extracts. Upon 5'-AMP-Sepharose chromatography MDA dehydrogenase was separated into two isoforms, I and II. Isoform I was eluted from the affinity carrier with a 0.1 M phosphate buffer pH 8.0. This isoform had a broad substrate specificity towards aliphatic and aromatic aldehydes. Kinetic studies showed that short- and medium-chain aliphatic aldehydes (C2-C6) were characterized by the lowest Km values and the highest Vmax values. The Km' values for MDA and acetaldehyde were 2.8 microM and 0.69 microM, respectively. Isoform II was eluted with a 0.1 M phosphate buffer pH 8.0 containing 0.5 mM NAD, was the most active with medium- and long-chain aliphatic aldehydes (C6-C11) and had Km values for MDA and acetaldehyde equal to 37 microM and 52 microM, respectively. Isoform I was much more sensitive towards disulfiram inhibition than isoform II. Both isoforms had an identical molecular mass (93 kD) upon gel filtration. It is concluded that MDA dehydrogenase isoform I is identical to mitochondrial aldehyde dehydrogenase having a low Km for acetaldehyde, whereas isoform II may be localized in liver cytosol. The role of aldehyde dehydrogenases in the metabolism of aldehydes derived from lipid peroxidation is discussed.  相似文献   

17.
We examined the effect of a novel disulfide bond engineered in subtilisin E from Bacillus subtilis based on the structure of a thermophilic subtilisin-type serine protease aqualysin I. Four sites (Ser163/Ser194, Lys170/Ser194, Lys170/Glu195, and Pro172/Glu195) in subtilisin E were chosen as candidates for Cys substitutions by site-directed mutagenesis. The Cys170/Cys195 mutant subtilisin formed a disulfide bond in B. subtilis, and showed a 5-10-fold increase in specific activity for an authentic peptide substrate for subtilisin, N-succinyl-L-Ala-L-Ala-L-Pro-L-Phe-p-nitroanilide, compared with the single-Cys mutants. However, the disulfide mutant had a 50% decrease in catalytic efficiency due to a smaller k(cat) and was thermolabile relative to the wild-type enzyme, whereas it was greatly stabilized relative to its reduced form. These results suggest that an electrostatic interaction between Lys170 and Glu195 is important for catalysis and stability in subtilisin E. Interestingly, the disulfide mutant was found to be more stable in polar organic solvents, such as dimethylformamide and ethanol, than the wild-type enzyme, even under reducing conditions; this is probably due to the substitution of uncharged Cys by charged surface residues (Lys170 and Glu195). Further, the amino-terminal engineered disulfide bond (Gly61Cys/Ser98Cys) and the mutation Ile31Leu were introduced to enhance the stability and catalytic activity. A prominent 3-4-fold increase in the catalytic efficiency occurred in the quintet mutant enzyme over the range of dimethylformamide concentration (up to 40%).  相似文献   

18.
Crotonaldehyde was oxidized by disrupted rat liver mitochondrial fractions or by intact mitochondria at rates that were only 10 to 15% that of acetaldehyde. Although a poor substrate for oxidation, crotonaldehyde is an effective inhibitor of the oxidation of acetaldehyde by mitochondrial aldehyde dehydrogenase, by intact mitochondria, and by isolated hepatocytes. Inhibition by crotonaldehyde was competitive with respect to acetaldehyde, and the Ki for crotonaldehyde was about 5 to 20 microM. Crotonaldehyde had no effect on the oxidation of glutamate or succinate. Very low levels of acetaldehyde were detected during the metabolism of ethanol. Crotonaldehyde increased the accumulation of acetaldehyde more than 10-fold, indicating that crotonaldehyde, besides inhibiting the oxidation of added acetaldehyde, also inhibited the oxidation of acetaldehyde generated by the metabolism of ethanol. Formaldehyde was a substrate for the low-Km mitochondrial aldehyde dehydrogenase, as well as for a cytosolic, glutathione-dependent formaldehyde dehydrogenase. Crotonaldehyde was a potent inhibitor of mitochondrial oxidation of formaldehyde, but had no effect on the activity of formaldehyde dehydrogenase. In hepatocytes, crotonaldehyde produced about 30 to 40% inhibition of formaldehyde oxidation, which was similar to the inhibition produced by cyanamide. This suggested that part of the formaldehyde oxidation occurred via the mitochondrial aldehyde dehydrogenase, and part via formaldehyde dehydrogenase. The fact that inhibition by crotonaldehyde is competitive may be of value since other commonly used inhibitors of aldehyde dehydrogenase are irreversible inhibitors of the enzyme.  相似文献   

19.
Toxic medium chain length alkanals, alkenals, and 4-hydroxyalkenals that are generated during lipid peroxidation are potential substrates for aldehyde dehydrogenase (ALDH) isoforms. We have developed transgenic cell lines to examine the potential for either human ALDH1A1 or ALDH3A1 to protect against damage mediated by these toxic aldehydes. Using crude cytosols from stably transfected cell lines, these aldehydes were confirmed to be excellent substrates for ALDH3A1, but were poorly oxidized by ALDH1A1. Expression of ALDH3A1 by stable transfection in V79 cells conferred a high level of protection against growth inhibition by the medium-chain length aldehyde substrates with highest substrate activity, including hexanal, trans-2-hexenal, trans-2-octenal, trans-2-nonenal, and 4-hydroxy-2-nonenal (HNE). This was reflected in a parallel ability of ALDH3A1 to prevent depletion of glutathione by these aldehydes. Expression of hALDH3 completely blocked the potent induction of apoptosis by HNE in both V79 cells and in a RAW 264.7 murine macrophage cell line, consistent with the observed total prevention of HNE-protein adduct formation. Structure–activity studies indicated that the rank order of potency for the contributions of HNE functional groups to toxicity was aldehyde ≥C2=C3 double bond>>C4-hydroxyl group. Oxidation of the aldehyde moiety of HNE to a carboxyl by ALDH3A1 expressed in stably transfected cell lines drastically reduced its potency for growth inhibition and apoptosis induction. In contrast, ALDH1A1 expression provided only moderate protection against trans-2-nonenal (t2NE), and none against the other six–nine carbon aldehydes. Neither ALDH1A1 nor ALDH3A1 conferred any protection against acrolein, acetaldehyde, or chloroacetaldehyde. A small degree of protection against malondialdehyde was afforded by ALDH1A1, but not ALDH3A1. Paradoxically, cells expressing ALDH3A1 were 1.5-fold more sensitive to benzaldehyde toxicity than control V79 cells. These studies demonstrate that expression of class 3 ALDH, but not class 1 ALDH, can be an important determinant of cellular resistance to toxicity mediated by aldehydes of intermediate chain length that are produced during lipid peroxidation.  相似文献   

20.
Crystal structures of two engineered thiol trypsins   总被引:3,自引:0,他引:3  
We have determined the three-dimensional structures of engineered rat trypsins which mimic the active sites of two classes of cysteine proteases. The catalytic serine was replaced with cysteine (S195C) to test the ability of sulfur to function as a nucleophile in a serine protease environment. This variant mimics the cysteine trypsin class of thiol proteases. An additional mutation of the active site aspartate to an asparagine (D102N) created the catalytic triad of the papain-type cysteine proteases. Rat trypsins S195C and D102N,S195C were solved to 2.5 and 2.0 A, respectively. The refined structures were analyzed to determine the structural basis for the 10(6)-fold loss of activity of trypsin S195C and the 10(8)-fold loss of activity of trypsin D102N,S195C, relative to rat trypsin. The active site thiols were found in a reduced state in contrast to the oxidized thiols found in previous thiol protease structures. These are the first reported structures of serine proteases with the catalytic centers of sulfhydryl proteases. Structure analysis revealed only subtle global changes in enzyme conformation. The substrate binding pocket is unaltered, and active site amino acid 102 forms hydrogen bonds to H57 and S214 as well as to the backbone amides of A56 and H57. In trypsin S195C, D102 is a hydrogen-bond acceptor for H57 which allows the other imidazole nitrogen to function as a base during catalysis. In trypsin D102N,S195C, the asparagine at position 102 is a hydrogen-bond donor to H57 which places a proton on the imidazole nitrogen proximal to the nucleophile. This tautomer of H57 is unable to act as a base in catalysis.(ABSTRACT TRUNCATED AT 250 WORDS)  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