Article

The use of pH-gradient ion-exchange chromatography to separate sheep liver cytoplasmic aldehyde dehydrogenase from mitochondrial enzyme contamination, and observations on the interaction between the pure cytoplasmic enzyme and disulfiram

Authors:
To read the full-text of this research, you can request a copy directly from the authors.

Abstract

1. Sheep liver cytoplasmic aldehyde dehydrogenase can be purified from contamination with the mitochondrial form of the enzyme by pH-gradient ion-exchange chromatography. The method is simple, reproducible and efficient. 2. The purified cytoplasmic enzyme retains about 2% of its original activity in the presence of a large excess of disulfiram. This suggests that the disulfiram-reactive thiol groups are not essential for covalent interaction with the aldehyde substrate during catalysis, as has sometimes been suggested. 3. Between 1.5 and 2.0 molecules of disulfiram per tetrameric enzyme molecule account for the observed loss of activity, suggesting that the enzyme may have only two functional active sites. 4. Experiments show that disulfiram-modified enzyme retains the ability to bind NAD+ and NADH.

No full-text available

Request Full-text Paper PDF

To read the full-text of this research,
you can request a copy directly from the authors.

Article
The published data concerning the generation of pH gradients in liquid chromatography are systematized. The pH gradients are classified according to the methods of their generation and the gradient profile; the application of gradients is considered. Particular attention is given to adsorbents and stationary phases for producing internal pH gradients in chromatofocusing. Prospects for using pH gradients in analytical chemistry are discussed.
Article
p-Hydroxyacetophenone was coupled to epoxy-activated Sepharose 6B to generate an affinity chromatographic matrix to purify aldehyde dehydrogenase. Purified beef liver mitochondrial aldehyde dehydrogenase specifically bound to the support and could be eluted with p-hydroxyacetophenone. A post-ammonium sulfate (30-55%) fraction of bovine liver was applied to the affinity gel column and aldehyde dehydrogenase was effectively purified, although not to complete homogeneity, indicating the potential selectivity of the matrix. Both beef liver cytosolic and mitochondrial aldehyde dehydrogenase bound to the column. A post-Cibacron blue Sepharose Cl-6B affinity-fractionated liver mitochondrial aldehyde dehydrogenase was purified to complete homogeneity by p-hydroxyacetophenone-Sepharose, thus eliminating the need for the isoelectric focusing step often employed. p-Hydroxyacetophenone was found to be a competitive inhibitor against propionaldehyde and noncompetitive against NAD. Escherichia coli lysates of recombinantly expressed aldehyde dehydrogenase were purified from E. coli lysates with one major 25-kDa protein contaminant also binding to the column, as detected by sodium dodecyl sulfate-polyacrylamide gel electrophoresis analysis. The 25-kDa contaminant was found to be chloramphenicol acetyl transferase from sequence analysis and binding studies.
Article
The effects of S-methyl diethyldithiocarbamate, S-methyl diethylmonothiocarbamate and bis(diethylcarbamoyl) disulphide on sheep liver cytoplasmic aldehyde dehydrogenase were investigated in vitro. The first compound has negligible effect. The second one is a weak inhibitor of the esterase activity of the enzyme and a weaker inhibitor of the dehydrogenase activity. A very low concentration of the third compound, however, acts as a potent inactivator of aldehyde dehydrogenase, similar in this respect to disulfiram, although somewhat slower to react. The possible involvement of these compounds in the physiological phenomenon known as the disulfiram ethanol reaction is discussed.
Article
Bromoacetophenone (2-bromo-1-phenylethanone) has been characterized as an affinity reagent for human aldehyde dehydrogenase (EC 1.2.1.3) [MacKerell, MacWright & Pietruszko (1986) Biochemistry 25, 5182-5189], and has been shown to react specifically with the Glu-268 residue [Abriola, Fields, Stein, MacKerell & Pietruszko (1987) Biochemistry 26, 5679-5684] with an apparent inactivation stoichiometry of two molecules of bromoacetophenone per molecule of enzyme. The specificity of bromoacetophenone for reaction with Glu-268, however, is not absolute, owing to the extreme reactivity of this reagent. When bromo[14C]acetophenone was used to label the human cytoplasmic E1 isoenzyme radioactively and tryptic fragmentation was carried out, peptides besides that containing Glu-268 were found to have reacted with reagent. These peptides were purified by h.p.l.c. and analysed by sequencing and scintillation counting to quantify radioactive label in the material from each cycle of sequencing. Reaction of bromoacetophenone with the aldehyde dehydrogenase molecule during enzyme activity loss occurs with two residues, Glu-268 and Cys-302. The activity loss, however, appears to be proportional to incorporation of label at Glu-268. The large part of incorporation of label at Cys-302 occurs after the activity loss is essentially complete. With both Glu-268 and Cys-302, however, the incorporation of label stops after one molecule of bromoacetophenone has reacted with each residue. Reaction with other residues continues after activity loss is complete.
Article
Sheep liver cytoplasmic aldehyde dehydrogenase was labelled by reaction with the substrate p-nitrophenyl di[14C]methylcarbamate. After tryptic digestion and peptide fractionation the labelled residue was identified as Cys-302. This is the first unequivocal identification of the essential enzymic nucleophile in the esterase activity of aldehyde dehydrogenase. By implication, Cys-302 is probably also the residue that is acylated by aldehyde substrates and the first residue that is modified by disulfiram.
Article
Although studies on the effect of pH on aldehyde dehydrogenase have been carried out in a number of laboratories, the results of these studies do not give a complete picture of the mechanism of the aldehyde dehydrogenase catalyzed oxidation of propionaldehyde over an extended pH range. The enzyme catalyzed oxidation reaction at pH 7.0 and 7.6 is generally agreed to occur by the following ordered mechanism: $$E \rightleftharpoons E.NA{D^ + } \rightleftharpoons E.NA{D^ + }.ALD \rightleftharpoons E.NADH.acyl \to *E.NADH \rightleftharpoons E.NADH \rightleftharpoons E$$Scheme I At least at low propionaldehyde concentrations, the rate of the reaction is controlled by the slow conformational change which occurs during the release of NADH from the enzyme (Blackwell et al., 1987; Dickinson, 1985).
Article
A nucleophilic group in the active site of aldehyde dehydrogenase, which covalently binds the aldehyde moiety during the enzyme-catalyzed oxidation of aldehydes to acids, was acylated with the chromophoric aldehyde trans-4-(N,N-dimethylamino)cinnamaldehyde (DACA). Acyl-enzyme trapped by precipitation with perchloric acid was digested with trypsin, and the peptide associated with the chromophoric group was isolated and shown to be Gln-Ala-Phe-Gln-Ile-Gly-Ser-Pro-Trp-Arg. After redigestion with thermolysin, the chromophore was associated with the C-terminal hexaresidue part. If the chromophore is attached to this peptide, serine would be expected to bind the aldehyde and lead to the required acylated derivative. Differential labeling experiments were performed in which all free thiol groups on the acylated enzyme were blocked by carboxymethylation. The acyl chromophore was then removed by controlled hydrolysis and the protein reacted with [14C]iodoacetamide. No 14C-labeled tryptic peptides were isolated, suggesting that the sulfur of a cysteine cannot be the acylated residue in the precipitated acyl-enzyme.
Article
Various kinetic approaches were carried out to investigate kinetic attributes for the dual coenzyme activities of mitochondrial aldehyde dehydrogenase from rat liver. The enzyme catalyses NAD(+)- and NADP(+)-dependent oxidations of ethanal by an ordered bi-bi mechanism with NAD(P)+ as the first reactant bound and NAD(P)H as the last product released. The two coenzymes presumably interact with the kinetically identical site. NAD+ forms the dynamic binary complex with the enzyme, while the enzyme-NAD(P)H complex formation is associated with conformation change(s). A stopped-flow burst of NAD(P)H formation, followed by a slower steady-state turnover, suggests that either the deacylation or the release of NAD(P)H is rate limiting. Although NADP+ is reduced by a faster burst rate, NAD+ is slightly favored as the coenzyme by virtue of its marginally faster turnover rate.
Article
The effect of various thiol-modifying reagents on the esterase activity of sheep liver cytoplasmic aldehyde dehydrogenase is reported here. Both symmetrical reagents (disulfiram, 2,2'- and 4,4'-dithiodipyridines) and unsymmetrical reagents (methyl diethylthiocarbamyl disulphide, methyl 2- and 4-pyridyl disulphides) were investigated. The results suggest that all the modifiers react to varying extents with a pair of enzymic thiol groups ('A' and 'B'), and that the more specifically group 'A' is modified, the more the enzyme is inactivated. This supports the idea that group 'A' may be the essential nucleophile in the reaction catalysed by aldehyde dehydrogenase. Modification of group 'B' may or may not reduce the esterase activity depending on the nature of the label introduced. The results of the present experiments and of previous similar experiments concerning the dehydrogenase activity of the enzyme are consistent with the proposal that a common active site is responsible for both esterase and dehydrogenase activities.
Article
1. The pig enzyme was purified to homogeneity and was found to be a tetramer of apparently identical subunits. 2. The pig enzyme was found to contain 1 mol NADH/mol enzyme which is tightly bound, which is not directly involved in catalysis and which so far has not been removed from the enzyme so as to produce an active apoenzyme. 3. The pig enzyme seems to contain only one functioning active site/tetramer. 4. The pig and sheep enzymes are compared in respect of NADH binding, substrate specificity, immunological response and surface charge.
Article
Cytoplasmic aldehyde dehydrogenase may be modified by reaction with p-nitrophenyl dimethylcarbamate to give a stable E-X-CO-NMe2 species that is an analogue of the usual labile acyl-enzyme involved in the enzyme's reactions. [X is derived from an enzymic nucleophilic group.] This species still contains the tightly bound NADH that is present in the native enzyme. When further NADH binds to E-X-CO-NMe2 its fluorescence is enhanced over 4 times more than when it binds to unmodified enzyme; this fluorescence is completely unaffected by high propionaldehyde concentration and only slightly affected by p-nitrobenzaldehyde. The modified species has 1.0 NADH binding site in the absence of Mg2+ and 1.67 sites in its presence. The rate of dissociation of E-X-CO-NMe2.NADH is biphasic (k 3.4 and 1.8 min-1) and is considerably lower than that of E.NADH; the presence of Mg2+ slows the process even more (k 0.47 and 0.37 min-1). The implications of these studies towards a greater understanding of the nature of aldehyde dehydrogenase-catalysed reactions are discussed.
Article
The effects of modifiers (NAD+, NADH, propionaldehyde, chloral hydrate, diethylstilboestrol and p-nitrobenzaldehyde) on the hydrolysis of p-nitrophenyl (PNP) pivalate (PNP trimethylacetate) catalysed by cytoplasmic aldehyde dehydrogenase are reported. In each case a different inhibition pattern is obtained to that observed when the substrate is PNP acetate; for example, propionaldehyde and chloral hydrate competitively inhibit the hydrolysis of PNP acetate, but are mixed inhibitors with PNP pivalate. The kinetic results can be rationalized in terms of different rate-determining steps: acylation of the enzyme in the case of the pivalate but acyl-enzyme hydrolysis for the acetate. This is confirmed by stopped-flow studies, in which a burst of p-nitrophenoxide is observed when the substrate is PNP acetate, but not when it is the pivalate. PNP pivalate inhibits the dehydrogenase activity of the enzyme competitively with the aldehyde substrate; this is most simply explained if the esterase and dehydrogenase reactions occur at a common enzymic site.
Article
Cephalosporin antibiotics with a 1-methyltetrazole-5-thio side chain have the ability to cause an unpleasant flushing reaction if they are taken some time before the drinking of alcohol. It is proposed that the explanation for this is that the side chain becomes liberated in vivo and oxidized to 5,5'-dithiobis(1-methyltetrazole) or to a mixed disulfide analogue which then inactivates aldehyde dehydrogenase. Support for this proposal is given by the results below concerning the interaction in vitro between the disulfides and sheep liver cytoplasmic aldehyde dehydrogenase. 5,5'-Dithiobis(1-methyltetrazole) has a rapid and pronounced inactivatory effect, very similar in many ways (though not identical) to that of disulfiram, to which it has a structural similarity. (Disulfiram is widely used therapeutically to deter alcoholics from drinking.) 1-Methyl-5-methylthiotetrazole (which is a simple model of the antibiotics) and the free 1-methyltetrazole-5-thiol have no effect on the enzyme in vitro, but methyl 5-(1-methyltetrazolyl) disulfide is a potent inactivator; this also supports the proposed pathway.
Article
Initial-rate measurements and stopped-flow spectrophotometric experiments over a wide range of pH implicate an enzyme group of pKa approximately 6.6 affecting the aldehyde binding reactions. It is possible, though not proved, that the group involved is the cysteine residue involved in catalysis. Stopped-flow fluorescence studies show that a group of pKa greater than 8.5 facilitates hydrolysis of the NADH-containing acyl-enzyme species. The identity of this group is quite unknown. Studies with 4-nitrobenzaldehyde show that this substrate gives marked substrate inhibition at quite low (less than 20 microM) concentrations. The mechanism of catalysis seems to be the same as for propionaldehyde oxidation. It is argued that proton release occurs with both substrates on hydrolysis of the NADH-containing acyl-enzyme and not before hydride transfer, as has been previously suggested [Bennett, Buckley & Blackwell (1982) Biochemistry 21, 4407-4413].
Article
Assays of UDP-glucose dehydrogenase at pH 6.0 show long (10-15 min) lag periods before the steady-state rate is established, but at pH 9.0 no lag is observed. At intermediate pH values the lag is progressively shorter as the pH becomes more alkaline. The behaviour of the enzyme in assays at neutral and acid pH depends on the pH and concentration of the enzyme used to initiate the assay. The steady-state rate at pH 6.0 is strongly concentration-dependent. It is suggested that these phenomena arise because of the slow dissociation of an inactive enzyme species to an active one. Purified preparations of the enzyme release approx. 1 mol of a UDP-sugar/mol of enzyme subunit on denaturation. The identity of the UDP-sugar is unknown.
Article
Stopped-flow spectrophotometric experiments show that modification by disulfiram not only lowers the steady-state rates but also decreases the size of bursts seen in both dehydrogenase and esterase reactions catalysed by sheep liver cytoplasmic aldehyde dehydrogenase. This observation is consistent with the proposal that a catalytically essential group is modified by disulfiram and that this group mediates both dehydrogenase and esterase activities.
Article
The activity of aldehyde dehydrogenase (ALDH, EC 1.2.1.3) was measured in different fractions of human blood. Of the recovered activity 99% was detected in the intracellular fraction of the erythrocytes. The results also indicated the presence of ALDH activity in the leukocytes, since an increased activity was obtained after cultivation of the cells in the presence of a mitogen. No activity was detected in platelets, plasma, or erythrocyte membranes. Nonlinear Lineweaver-Burk plots were obtained with acetaldehyde, 3,4-dihydroxyphenylacetaldehyde, and indole-3-acetaldehyde as the substrates. The apparent Km values, calculated from the low and high substrate concentration ranges of the curves, were much lower for 3,4-dihydroxyphenylacetaldehyde and indole-3-acetaldehyde than for acetaldehyde. Disulfiram caused almost complete inhibition of the blood ALDH activity in assays with acetaldehyde as the substrate, whereas 15-30% of the activity remained unaffected in assays with 3,4-dihydroxyphenylacetaldehyde and indole-3-acetaldehyde. Kinetic experiments using the mixed substrate method and isoelectric focusing of a partially purified sample of blood did not reveal the presence of more than one isozyme.
Article
Full-text available
The inhibition of mitochondrial (pI 5) horse liver aldehyde dehydrogenase by disulfiram (tetraethylthiuram disulphide) was investigated to determine if the drug was an active-site-directed inhibitor. Stoichiometry of inhibition was determined by using an analogue, [35S]tetramethylthiuram disulphide. A 50% loss of the dehydrogenase activity was observed when only one site per tetrameric enzyme was modified, and complete inactivation was not obtained even after seven sites per tetramer were modified. Modification of only two sites accounted for a loss of 75% of the initial catalytic activity. The number of functioning active sites per tetrameric enzyme, as determined by the magnitude of the pre-steady-state burst of NADH formation, did not decrease until approx. 75% of the catalytic activity was lost. These data indicate that disulfiram does not modify the essential nucleophilic amino acid at the active site of the enzyme. The data support an inactivation mechanism involving the formation of a mixed disulphide with a non-essential cysteine residue, resulting in a lowered specific activity of the enzyme.
Article
A review is made of the pharmacological, biochemical and chemical aspects of the unpleasant 'Antabuse-like' reaction that may be induced in drinkers of alcohol by pre-treatment with certain beta-lactam antibiotics with a 1-methyltetrazole-5-thiol sidechain (such as moxalactam, cefamandole and cefoperazone). The symptoms are due to abnormally elevated blood acetaldehyde levels consequent upon the inactivation of hepatic aldehyde dehydrogenase. There is very little direct effect of the antibiotics on this enzyme and therefore it is concluded that a reactive metabolite of the antibiotics' essential sidechain is responsible for the reaction. A likely candidate for this active species is either the symmetrical disulphide 5,5'-dithiobis(1-methyltetrazole) formed by oxidation of 1-methyltetrazole-5-thiol, or the related mixed disulphide, methyl 5-(1-methyltetrazolyl) disulphide. The first of these is a potent inactivator of cytoplasmic aldehyde dehydrogenase only, the second affects both cytoplasmic and mitochondrial isoenzymes. 1-Methyltetrazole-5-thiol or derivatives have the potential to be used therapeutically as 'anti-alcohol' compounds in the same way as disulfiram (Antabuse) or calcium cyanamide.
Article
Stopped-flow experiments in spectrophotometric and fluorescence modes reveal different aspects of the aldehyde dehydrogenase mechanism. Spectrophotometric experiments show a rapid burst of NADH production whose course is not affected by Mg2+. The slower burst seen in the fluorescence mode is markedly accelerated by Mg2+. It is argued that the fluorescence burst accompanies acyl-enzyme hydrolysis and, therefore, that Mg2+ increases the rate of this process. Experiments on the hydrolysis of p-nitrophenyl propionate indicate that acyl-enzyme hydrolysis is indeed accelerated by Mg2+ and a combination of Mg2+ and NADH. Vmax. values for p-nitrophenyl propionate hydrolysis in the presence of NADH and NADH and Mg2+ agree closely with the specific rates of acyl hydrolysis from the E . NADH . acyl and E . NADH . acyl . Mg2+ complexes seen in the dehydrogenase reaction with propionaldehyde. These observations support the view that esterase and dehydrogenase activities occur at the same site on the enzyme. Other evidence is presented to support this conclusion.
Article
The binding of diethylstilbestrol (DES) to aldehyde dehydrogenase (ALDH) has a very similar effect on the dehydrogenase activity of the enzyme as has modification of the enzyme by 2,2'-dithiodipyridine [Kitson, T.M. (1982) Biochem. J. 207, 81-89]. The latter modification may occur at the site of the esterase activity of the enzyme [Kitson, T.M. (1985) Biochem. J. 228, 765-767]. This suggests that DES might be a competitive inhibitor of the esterase reaction. However, in the absence of oxidized nicotinamide adenine dinucleotide (NAD+) or reduced nicotinamide adenine dinucleotide (NADH), and at low concentrations of substrate (4-nitrophenyl acetate, PNPA), DES is a potent partial noncompetitive inhibitor. It is concluded therefore that DES binds at a site different from the esterase active site and that the enzyme-DES complex retains some ability to act as an esterase. High concentrations of PNPA appear to displace DES from its binding site. In the presence of NAD+, DES is a weaker inhibitor, and in the presence of NADH, DES has very little effect. Esterase activity is enhanced by NADH when PNPA concentrations are high but is inhibited when they are low. The rate of reaction of ALDH with 2,2'-dithiodipyridine is only slightly reduced by DES, suggesting that the site at which thiol modifiers react and the DES binding site are different. When ALDH is modified by 2,2'-dithiodipyridine, it has reduced esterase activity, which declines further as the modified enzyme loses its 2-thiopyridyl label. In the presence of NAD+, chloral hydrate is a simple competitive inhibitor of the esterase reaction.(ABSTRACT TRUNCATED AT 250 WORDS)
Article
Full-text available
Incubation of sheep liver cytoplasmic aldehyde dehydrogenase with the substrate 4-nitrophenyl [14C]acetate in the presence of NADH leads to the formation of 14C-labelled acetaldehyde. This observation strongly supports the idea that the esterase and dehydrogenase activities of the enzyme occur at the same site and involve the intermediacy of a common acyl-enzyme.
Article
The dissociation of the aldehyde dehydrogenase X NADH complex was studied by displacement with NAD+. The association reaction of enzyme and NADH was also studied. These processes are biphasic, as shown by McGibbon, Buckley & Blackwell [(1977) Biochem. J. 165, 455-462], but the details of the dissociation reaction are significantly different from those given by those authors. Spectral and kinetic experiments provide evidence for the formation of abortive complexes of the type enzyme X NADH X aldehyde. Kinetic studies at different wavelengths with transcinnamaldehyde as substrate provide evidence for the formation of an enzyme X NADH X cinnamoyl complex. Hydrolysis of the thioester relieves a severe quenching effect on the fluorescence of enzyme-bound NADH.
Article
It is proposed that cytoplasmic aldehyde dehydrogenase possesses a pair of important reactive thiol groups, A and B. Group A is labelled by disulfiram and the enzyme is inactivated; subsequently group B displaces the dithiocarbamate label and an enzymic disulphide is formed. On the other hand, it appears that group B is labelled by 2,2'-dithiodipyridine resulting in activation of the enzyme. Again, the label (2-thiopyridone) is later displaced, this time presumably by group A, giving rise to loss of enzymic activity and formation of the same disulphide species as is produced by disulfiram. Methyl diethylthiocarbamyl disulphide and methyl 2-pyridyl disulphide supply the same label (MeS-) but the first compound inactivates the enzyme while the second activates it. It is concluded that the first of these reagents modifies group A and the second group B. It appears that methyl 4-pyridyl disulphide may react non-specifically with both groups A and B. Group A is a possible candidate for a catalytically essential nucleophile in the actions of aldehyde dehydrogenase.
Article
High concentrations of aldehydes slow the inactivation of cytoplasmic aldehyde dehydrogenase by disulfiram and also slow the reaction of the enzyme with 2,2'-dithiodipyridine. It is concluded that a low-affinity aldehyde-binding site is probably the site at which thiol-group modifiers react with aldehyde dehydrogenase, as well as being the active site for hydrolysis of 4-nitrophenyl acetate.
Article
Despite the fact that it is an aldehyde, glyoxylic acid is not a substrate for sheep liver cytoplasmic aldehyde dehydrogenase; instead it functions as an inhibitor of both the esterase and dehydrogenase activities. From a consideration of the inhibition patterns it is concluded that glyoxylic acid does not bind in the catalytic propionaldehyde-binding domain, thus confirming the two-site model as proposed previously. Since the corresponding neutral methyl ester is a substrate it is suggested that the catalytic binding domain must contain a negatively charged group which prevents the binding of glyoxylic acid. Steady-state and pre-steady-state kinetic studies indicate that glyoxylic acid inhibits the dehydrogenase activity by converting the enzyme into a dead-end form which cannot undergo the catalytically essential conformational change. Incubation of the enzyme with NAD+ and glyoxylic acid for 10 min before the addition of propionaldehyde gave rise to hysteresis effects which can be explained on the basis of a slow isomerization of the enzyme X NAD+ X glyoxylic acid complex.
Article
A transient release of protons with an amplitude corresponding to one proton per active site has been observed for the oxidation of propionaldehyde, acetaldehyde, and benzaldehyde by sheep liver cytoplasmic aldehyde dehydrogenase at pH 7.6 with phenol red as indicator. At saturating substrate levels, the rate constants for the proton burst are in each case the same, and for acetaldehyde and propionaldehyde show the same dependence on the concentrations of the substrates, as the rate constants for the transient production of NADH reported previously [MacGibbon, A.K.H., Blackwell, L.F., & Buckley, P.D. (1977) Biochem. J. 167, 469-477]. Although, with propionaldehyde as a substrate, a full proton burst is also observed at pH 6.0, no proton burst is observed at pH 9.0. For 4-nitrobenzaldehyde, there is no burst in NADH production, but a burst in proton release is observed, showing that proton release precedes hydride transfer. No protons were released during the binding of the substrate analogues acetone and chloral hydrate nor on reaction of the enzyme with the inhibitor tetraethylthiuram disulfide (disulfiram). A model is proposed in which the rate-limiting step in the pre-steady-state phase of the reaction is a conformational change which occurs after the binding of aldehydes to the enzyme. As a result of the conformational change, the environment of a functional group on the enzyme, which initially has a pKa of about 8.5, is perturbed to give a final pKa value for the group of less than 5. Computer simulations were used to show that the model accurately reproduces all of the experimental data. The lack of observation of a second transient proton release, as required by the overall stoichiometry, argues that its release occurs in a slow step prior to NADH dissociation.
Article
The binding of NADH and NAD+ by cytoplasmic aldehyde dehydrogenase was studied by various direct and indirect methods. At pH 7.0 at 25 degrees C there appears to be approx. 1 binding site for both nucleotides per 200 000 daltons of protein, although the NAD+-binding results are rather uncertain. Estimates of the dissociation constants of the E . NADH and E . NAD+ complexes under the stated conditions are also presented. Preparations of enzyme are sometimes found to contain significant amounts of very tightly bound NAD+ and NADH. The implications of these findings are discussed.
Article
The pI approximately 5.2 isoenzymes of mitochondrial aldehyde dehydrogenase were separated from the other isoenzymes by pH-gradient chromatography on DEAE-Sephacel. The pI approximately 5.2 material is immunologically identical with cytosolic aldehyde dehydrogenase. It also shows sensitivity to 20 microM-disulfiram and insensitivity to 4M-urea in assays. These and other criteria seem to establish that the material is identical with the cytosolic enzyme. Mitochondrial enzyme that had been purified to remove pI approximately 5.2 isoenzymes shows concentration-dependent lag phases in assays. These effects are possibly due to the slow establishment of equilibrium between tetramer and either dimers or monomers, with the dissociated species being intrinsically more active than the tetramer.
Article
The molar absorptivity and A 1cm1% values for 150 proteins are reported. The conditions under which these values were obtained and citations to the sources of the data are also provided.
Article
2,2'-Dithiodipyridine reacts rapidly with sheep liver cytoplasmic aldehyde dehydrogenase in the presence of NAD +, resulting in activation of the enzyme by 2 to 2.5-fold (when assayed in the usual way). This is followed by the slow loss of most of the enzyme activity during the next few hours at 25 degrees C. 2-Thiopyridone is displaced from the labeled enzyme at approximately the same rate as activity is lost. This is explained in terms of the initial modification of an enzymatic thiol group (giving activation) followed by the reaction of the labeled group with a second enzymatic thiol group, resulting in the formation of a disulfide bond and the inactivation of the enzyme. 4,4'-Dithiodipyridine reacts in a broadly similar way, although both the loss of label and loss of activity are faster and do not correlate with each other as well as for the 2,2' isomer. The results suggest that the dithiodipyridines act to produce the same enzymatic disulfide bond as has been shown to arise from the reaction of the enzyme with disulfiram (a drug used in alcoholism treatment). The implications of the results are discussed with reference to the proposed mechanism of action of aldehyde dehydrogenase. It is concluded that the thiol group initially modified by disulfiram is unlikely to be catalytically essential to the dehydrogenase action of the enzyme.
Article
In the absence of NAD+, up to 12 SH groups on aldehyde dehydrogenase (ALDH) reacted rapidly with p-(chloromercuri)benzoate (PCMB); a slow reaction with more than twice this number of SH groups then occurred. When PCMB was added to an assay mixture at low (less than 100 microM) concentrations of propionaldehyde, the steady-state rate of production of NADH increased with increasing PCMB concentration up to a maximum activity at a [PCMB]/[ALDH] ratio of 1.9 and then decreased as the [PCMB]/[ALDH] ratio increased further. Under some conditions, activation, or inhibition, showed hysteretic effects as the initial slope after mixing changed to a final linear steady state in a first-order manner, the rate constants for which were proportional to the concentration of free PCMB. Activating levels of PCMB had little effect on the NADH and proton burst amplitudes or rate constants and did not affect the rate of dissociation or association of NADH. However, when a 20-fold excess of PCMB concentration over enzyme concentration was premixed with the enzyme, neither a burst nor a steady-state turnover of substrate was observed. It is concluded that activation arises from the tight binding of PCMB with a single thiol group per subunit which is exposed after the binding of NAD+ to the enzyme, followed by a slow conformational change which causes activation by altering the steady-state mechanism so that NADH dissociation becomes largely rate limiting.(ABSTRACT TRUNCATED AT 250 WORDS)
Article
Stoicheiometric amounts of [14C]disulfiram react rapidly with sheep liver cytoplasmic aldehyde dehydrogenase to give loss of catalytic activity and incorporation of the expected amount of radioactivity. In a subsequent slower reaction the label is lost from the enzyme without re-emergence of enzymic activity. The results imply that in vivo disulfiram may act as an oxidation-reduction catalyst for the inactivation of aldehyde dehydrogenase.
Article
Full-text available
Bovine lens cytoplasmic aldehyde dehydrogenase exhibits Michaelis-Menten kinetics with acetaldehyde, glyceraldehyde 3-phosphate, p-nitrobenzaldehyde, propionaldehyde, glycolaldehyde, glyceraldehyde, phenylacetylaldehyde and succinic semialdehyde as substrates. The enzyme was also active with malondialdehyde, and exhibited an esterase activity. Steady-state kinetic analyses show that the enzyme exhibits a compulsory-ordered ternary-complex mechanism with NAD+ binding before acetaldehyde. The enzyme was inhibited by disulfiram and by p-chloromercuribenzoate, and studies with with mercaptans indicated the involvement of thiol groups in catalysis.
Article
The aldehyde specificity of xanthine oxidase (xanthine:oxygen oxidoreductase, EC 1.2.3.2) has been reinvestigated. The biogenic aldehydes and succinate semialdehyde are reasonable substrates for xanthine oxidase. Pyrophosphate, which binds to xanthine oxidase, does not seem to affect significantly the enzyme's catalytic activity. The steady-state parameters for the oxidation of several substrates by xanthine oxidase and oxygen have been determined. Formaldehyde differs from xanthine and other aldehydes in phi 2, the parameter describing the reaction with oxygen. Substrate inhibition has been studied at high concentrations of xanthine with oxygen as the electron acceptor. The inhibition is hyperbolic and uncompetitive with respect to oxygen. This is possibly due to rate-limiting product release from molybdenum(IV) being slower than from molybdenum(VI).
Article
NAD-dependent aldehyde dehydrogenases (EC 1.2.1.3) were isolated from various subcellular organelles as well as from different regions of rat brain. The mitochondrial, microsomal, and cytosolic fractions were found to contain 40%, 28%, and 12%, respectively, of the total aldehyde dehydrogenase (5.28 +/- 0.44 nmol NADH/min/g tissue) found in rat brain homogenate when assayed with 70 muM propionaldehyde at pH 7.5. The total activity increased to 17.3 +/- 2.7 nmol NADH/min/g tissue when assayed with 5 mM propionaldehyde. Under these conditions the three organelles contained 49%, 23%, and 9%, respectively, of the activity. The enzyme isolated from cytosol possessed the lowest Km. The molecular weight of the enzyme isolated from all three subcellular organelles was approximately 100,000. Four activity bands were found by electrophoresis of crude homogenates, isolated mitochondria, or microsomes on cellulose acetate strips. Cytosol possessed just two of the forms. The total activity was essentially the same in homogenates obtained from cortex, subcortex, pons-medulla, or cerebellum. Further, the enzyme had the same molecular distribution and total activity in each of these four brain regions. Disulfiram was found to be an in vivo and in vitro inhibitor of the enzymes obtained from these brain regions. Mercaptoethanol, required for the stability of the enzyme, reversed the inhibition produced by disulfiram. The effect was greater for enzyme isolated from cytosol than from mitochondria. Calculations led to the prediction that aldehydes such as acetaldehyde are oxidized in cytosol.
Article
Magnesium chloride caused inhibition of the dehydrogenase activity of sheep liver cytoplasmic aldehyde dehydrogenase at all concentrations between pH 6 and 8 with no increase in the number of functioning subunits. There was also no decrease in the molecular weight as determined by gel filtration and laser light scattering experiments, results which are markedly different from those reported for the horse liver mitochondrial aldehyde dehydrogenase. There were changes in the spectroscopic and fluorescence properties of the enzyme, and enzyme-bound NADH, in the presence of magnesium ions. Steady-state inhibition studies revealed that magnesium ions exerted their inhibitory effect by decreasing Vmax for the reaction by binding to a metal ion binding site which was distinct from the coenzyme and substrate binding sites. The biphasic nature of the Lineweaver-Burk plots at high (millimolar) concentrations of propionaldehyde was shown to be consistent with a steady-state model in which two binding sites (a catalytic low-Km binding site and a high-Km modifier binding site) for propionaldehyde exist. Pre-steady-state kinetic studies showed that MgCl2 had no effect on the rates of NAD+ or NADH binding or on the rate constants for the bursts in production of NADH or proton release. However, the dissociation constants for E·NAD+ and E·NADH were significantly decreased in the presence of MgCl2, and the rate constants for dissociation of the coenzymes were shown to be decreased. At high concentrations of propionaldehyde, the inhibitory effect of MgCl2 could be almost entirely attributed to the tighter binding of NADH, but at low propionaldehyde concentrations, and for aromatic aldehydes, a more complex mechanism of inhibition must exist since the magnitude of the reduced kcat values was almost an order of magnitude less than the reduced value of the decay constant for the slow step of the NADH displacement process.
Article
1. Pre-modification of cytoplasmic aldehyde dehydrogenase by disulfiram results in the same extent of inactivation when the enzyme is subsequently assayed as a dehydrogenase or as an esterase. 2. 4-Nitrophenyl acetate protects the enzyme against inactivation by disulfiram, particularly well in the absence of NAD+. Some protection is also provided by chloral hydrate and indol-3-ylacetaldehyde (in the absence of NAD+). 3. When disulfiram is prevented from reacting at its usual site by the presence of 4-nitrophenyl acetate, it reacts elsewhere on the enzyme molecule without causing inactivation. 4. Enzyme in the presence of aldehyde and NAD+ is not at all protected against disulfiram. It is proposed that, under these circumstances, disulfiram reacts with the enzyme-NADH complex formed in the enzyme-catalysed reaction. 5. Modification by disulfiram results in a decrease in the amplitude of the burst of NADH formation during the dehydrogenase reaction, as well as a decrease in the steady-state rate. 6. 2,2'-Dithiodipyridine reacts with the enzyme both in the absence and presence of NAD+. Under the former circumstances the activity of the enzyme is little affected, but when the reaction is conducted in the presence of NAD+ the enzyme is activated by approximately 2-fold and is then relatively insensitive to the inactivatory effect of disulfiram. 7. Enzyme activated by 2,2'-dithiodipyridine loses most of its activity when stored over a period of a few days at 4 degrees C, or within 30 min when treated with sodium diethyldithiocarbamate. 8. Points for and against the proposal that the disulfiram-sensitive groups are catalytically essential are discussed.
Article
The kinetic properties of highly purified preparations of sheep liver cytoplasmic aldehyde dehydrogenase (preparations that had been shown to be free from contamination with the corresponding mitochondrial enzyme) were investigated with both propionaldehyde and butyraldehyde as substrates. At low aldehyde concentrations, double-reciprocal plots with aldehyde as the variable substrate are linear, and the mechanism appears to be ordered, with NAD+ as the first substrate to bind. Stopped-flow experiments following absorbance and fluorescence changes show bursts of NADH production in the pre-steady state, but the observed course of reaction depends on the pre-mixing conditions. Pre-mixing enzyme with NAD+ activates the enzyme in the pre-steady state and we suggest that the reaction mechanism may involve isomeric enzyme--NAD+ complexes. High concentrations of aldehyde in steady-state experiments produce significant activation (about 3-fold) at high concentrations of NAD+, but inhibition at low concentrations of NAD+. Such behaviour may be explained by postulating the participation of an abortive complex in product release. Stopped-flow measurements at high aldehyde concentrations indicate that the mechanism of reaction under these conditions is complex.
Article
Sheep liver cytoplasmic aldehyde dehydrogenase is strongly inhibited by Mg2+, Ca2+ and Mn2+. The inhibition is only partial, however, with 8-15% of activity remaining at high concentrations of these agents. In 50 mM-Tris/Hcl, pH 7.5, the concentrations giving half-maximal effect were: Mg2+, 6.5 micrometers; Ca2+, 15.2 micrometers; Mn2+, 1.5 micrometer. The esterase activity of the enzyme is not affected by such low metal ion concentrations, but appears to be activated by high concentrations. Fluorescence-titration and stopped-flow experiments provide evidence for interaction of Mg2+ with NADH complexes of the enzyme. As no evidence for the presence of increased concentrations of functioning active centres was obtained in the presence of Mg2+, it is concluded that effects of Mg2+ (and presumably Ca2+ and Mn2+ also) are brought about by trapping increased concentrations of NADH in a Mg2+-containing complex. This complex must liberate products more slowly than any of the complexes involved in the non-inhibited mechanism.
Article
1. The activation of sheep liver cytoplasmic aldehyde dehydrogenase by diethylstilboestrol and by 2,2'-dithiodipyridine is described. The effects of the two modifiers are very similar with respect to variation with acetaldehyde concentration, pH and temperature. Thus the degree of activation is maximal when the enzyme is assayed at approx. 1 mM-acetaldehyde, is greater at 25 degrees C than at 37 degrees C, and is greater at pH 7.4 than at pH 9.75. With low concentrations of acetaldehyde both modifiers decrease the enzyme activity. 2. Diethylstilboestrol affects the sheep liver cytoplasmic enzyme in a very similar way to that previously described for a rabbit liver cytoplasmic enzyme. Preliminary experiments show that the same is true for a preparation of human liver aldehyde dehydrogenase. It is proposed that sensitivity to diethylstilboestrol (and steroids) is a common property of all mammalian cytoplasmic aldehyde dehydrogenases.
Article
Human Class 1 and Class 2 aldehyde dehydrogenases have been sequenced at both the protein (Hempel et al., 1984, 1985) and DNA level (Hsu et al., 1988, 1989). Studies on the tertiary structure of aldehyde dehydrogenase are in progress (Baker et al., 1995, sheep Class 1; Hurley and Weiner, 1992, beef Class 2), but are not sufficiently advanced to suggest which amino acid residues are important in catalysis. Cys 302 is the only completely conserved cysteine in all known forms of the enzyme (Hempel et al., 1993), and labelling by various substrates and substrate analogues (von Bahr-Lindstrom et al., 1985; Kitson et al., 1991; Pietruszko et al., 1993) has implicated this residue as the probable active site nucleophile. This has been confirmed for the Class 2 enzyme by site-directed mutagenesis (Weiner et al., 1991). In order to establish whether Cys 302 is also the active site nucleophile for Class 1 aldehyde dehydrogenase we decided to carry out mutagenesis at Cys 302. A separate mutant was constructed in which Cys 301, the adjacent residue, was changed to alanine while Cys 302 was left unchanged.
Article
The kinetics of the cytosolic aldehyde dehydrogenase (ALDH) catalysed oxidation of propionaldehyde are complex. For example, at concentrations of propionaldehyde below about 200 μM propionaldehyde (low propionaldehyde concentration) the Lineweaver-Burk plot is linear, but at higher propionaldehyde concentration (20 mM) substrate activation occurs (MacGibbon et al., 1977a). At low propionaldehyde concentrations it is generally agreed that aldehyde dehydrogenase oxidises propionaldehyde by an ordered bi bi mechanism (Hill et al., 1991) as shown in scheme I in which isomerisation of binary enzyme.NADH complexes (k5) and subsequent NADH release (k6) are together rate limiting in the steady-state phase of the reaction. In scheme I *E represents a conformationally-rearranged form of the enzyme.
Article
Since 1948, disulfiram (Et2NCS-SS-CSNEt2) has been widely used in the treatment of alcoholism. This is because it lowers the activity of aldehyde dehydrogenase (AldDH) in vivo such that if alcohol is drunk, a build-up of acetaldehyde ensues, resulting in an unpleasant reaction (Kitson, 1977).
Article
To provide a molecular basis for understanding the possible mechanism of action of antidipsotropic agents in laboratory animals, aldehyde dehydrogenase (ALDH) isozymes were purified and characterized from the livers of hamsters and rats and compared with those from humans. The mitochondrial ALDHs from these species exhibit virtually identical kinetic properties in the oxidation and hydrolysis reactions. However, the cytosolic ALDH of human origin differs significantly from those of the rodents. Thus, for human ALDH-1, the Km value for acetaldehyde is 180 +/- 10 micromolar, whereas those for hamster ALDH-1 and rat ALDH-1 are 12 +/- 3 and 15 +/- 3 micromolar, respectively. Km values determined at pH 9.5 are virtually identical to those measured at pH 7.5. In vitro human ALDH-1 is 10 times less sensitive to disulfiram inhibition than are the hamster and rat cytosolic ALDHs. Competition between acetaldehyde and aromatic aldehydes or naphthaldehydes for the binding and catalytic sites of ALDHs shows their topography to be complex with more than one binding site. This also follows from data on substrate inhibition and activation, effects of NAD+ on ALDH-catalyzed hydrolysis of p-nitrophenyl esters, substrate specificity toward aldehydes and p-nitrophenyl esters, and inhibition by disulfiram in relation to oxidation and hydrolysis catalyzed by the ALDHs. The data further suggest that acetaldehyde cannot be considered as a "standard" ALDH substrate for studies aimed at aromatic ALDH substrates, e.g. biogenic aldehydes. Apparently, in human liver, only mitochondrial ALDH oxidizes acetaldehyde at physiological concentrations, whereas in hamster or rat liver, both the mitochondrial and cytosolic isozymes will do so.
Article
A purification procedure has been developed for the cytosolic aldehyde dehydrogenase of Saccharomyces cerevisiae that yields homogeneous enzyme. The enzyme seems to be a tetramer of identical 58 kDa subunits. The enzyme reaction is strongly stimulated by Mg2+ at low NADP+ concentrations but there is no absolute requirement for bivalent cations. The kinetics of the reaction have been studied in the presence and absence of MgCl2. NADP+ binding studies of the quenching of protein fluorescence in the presence and absence of MgCl2 show that the effect of Mg2+ is to increase the affinity of the enzyme for NADP+ by approx. 100-fold. NADP+ binding causes a slow conformational change in the enzyme and converts the enzyme from the inactive or low-activity form in which it is isolated into the fully active form. This conformational change seems to explain the marked lag-phases seen in enzyme assays. The enzyme is strongly inhibited by disulfiram and pyridoxal 5-phosphate.
Article
Evidence for dissociation of the tetrameric sheep liver cytosolic aldehyde dehydrogenase at pH 7.6 was reported by Blackwell et al. (1987). Subsequently the existence of order of mixing effects led Buckley et al. (1991) to propose that the tetrameric enzyme was dissociating into an inactive dimer or monomer. Gel filtration studies at pH 7.4 and at pH 5.2 confirmed that dissociation did take place significantly at pH 5.2, with the extent of the dissociation being concentration dependent (Buckley et al., 1991).
Article
1. The effect of disulfiram on the activity of the cytoplasmic and mitochondrial aldehyde dehydrogenases of sheep liver was studied. 2. Disulfiram causes an immediate inhibition of the enzyme reaction. The effect on the cytoplasmic enzyme is much greater than on the mitochondrial enzyme. 3. In both cases, the initial partial inhibition is followed by a gradual irreversible loss of activity. 4. The pH-rate profile of the inactivation of the mitochondrial enzyme by disulfiram and the pH-dependence of the maximum velocity of the enzyme-catalysed reaction are both consistent with the involvement of a thiol group. 5. Excess of 2-mercaptoethanol or GSH abolishes the effect of disulfiram. However, equimolar amounts of either of these reagents and disulfiram cause an effect greater than does disulfiram alone. It was shown that the mixed disulphide, Et2N-CS-SS-CH2-CH2OH, strongly inhibits aldehyde dehydrogenase. 6. The inhibitory effect of diethyldithiocarbamate in vitro is due mainly to contamination by disulfiram.
Article
The displacement of NADH from cytoplasmic aldehyde dehydrogenase (EC 1.2.1.3) from sheep liver was studied by using NAD+, 1,10-phenanthroline, ADP-ribose, deamino-NAD+ and pyridine-3-aldehyde-adenine dinucleotide as displacing agents, by following the decrease in fluorescence as a function of time. The data obtained could be fitted by assuming two first-order processes were occurring, a faster process with an apparent rate constant of 0.85 +/- 0.20 s-1 and a relative amplitude of 60 +/- 10% and a slower process with an apparent rate constant of 0.20 +/- 0.05 s-1 and a relative amplitude of 40 +/- 10% (except for pyridine-3-aldehyde-adenine dinucleotide, where the apparent rate constant for the slow process was 0.05 s-1). The displacement rates did not change significantly when the pH was varied from 6.0 to 9.0. Kinetic data are also reported for the dependence of the rate of binding of NADH to the enzyme on the total concentration of NADH. Detailed arguments are presented based on the isolation and purification procedures, the equilibrium coenzyme-binding studies and the kinetic data, which lead to the following model for the release of NADH from the enzyme: (formula: see article). The parameters that best fit the data are: k + 1 = 0.2 s-1; k - 1 = 0.05 s-1; k + 2 = 0.8 s-1 and k - 2 = 5 X 10(5)litre-mol-1-s-1. The slow phase of the NADH release is similar to the steady-state turnover number for substrates such as acetaldehyde and propionaldehyde and appears to contribute significantly to the limitation of the steady-state rate.
Article
Stopped-flow experiments in which sheep liver cytoplasmic aldehyde dehydrogenase (EC 1.2.1.3) was rapidly mixed with NAD(+) and aldehyde showed a burst of NADH formation, followed by a slower steady-state turnover. The kinetic data obtained when the relative concentrations and orders of mixing of NAD(+) and propionaldehyde with the enzyme were varied were fitted to the following mechanism: [Formula: see text] where the release of NADH is slow. By monitoring the quenching of protein fluorescence on the binding of NAD(+), estimates of 2x10(5) litre.mol(-1).s(-1) and 2s(-1) were obtained for k(+1) and k(-1) respectively. Although k(+3) could be determined from the dependence of the burst rate constant on the concentration of propionaldehyde to be 11s(-1), k(+2) and k(-2) could not be determined uniquely, but could be related by the equation: (k(-2)+k(+3))/k(+2) =50x10(-6)mol.litre(-1). No significant isotope effect was observed when [1-(2)H]propionaldehyde was used as substrate. The burst rate constant was pH-dependent, with the greatest rate constants occurring at high pH. Similar data were obtained by using acetaldehyde, where for this substrate (k(-2)+k(+3))/k(+2)=2.3x10 (-3)mol.litre(-1) and k(+3) is 23s(-1). When [1,2,2,2-(2)H]acetaldehyde was used, no isotope effect was observed on k(+3), but there was a significant effect on k(+2) and k(-2). A burst of NADH production has also been observed with furfuraldehyde, trans-4-(NN-dimethylamino)cinnamaldehyde, formaldehyde, benzaldehyde, 4-(imidazol-2-ylazo)benzaldehyde, p-methoxybenzaldehyde and p-methylbenzaldehyde as substrates, but not with p-nitrobenzaldehyde.
Article
Aldehyde dehydrogenase from sheep liver mitochondria was purified to homogeneity as judged by electrophoresis on polyacrylamide gels, and by sedimentation-equilibrium experiments in the analytical ultracentrifuge. The enzyme has a molecular weight of 198000 and a subunit size of 48000, indicating that the molecule is a tetramer. Fluorescence and spectrophotometric titrations indicate that each subunit can bind 1 molecule of NADH. Enzymic activity is completely blocked by reaction of 4mol of 5,5'-dithiobis-(2-nitrobenzoate)/mol of enzyme. Excess of disulfiram or iodoacetamide decreases activity to only 50% of the control value, and only two thiol groups per molecule are apparently modified by these reagents.
Article
The kinetics of the NAD+-dependent oxidation of aldehydes, catalysed by aldehyde dehydrogenase purified from sheep liver mitochondria, were studied in detail. Lag phases were observed in the assays, the length of which were dependent on the enzyme concentration. The measured rates after the lag phase was over were directly proportional to the enzyme concentration. If enzyme was preincubated with NAD+, the lag phase was eliminated. Double-reciprocal plots with aldehyde as the variable substrate were non-linear, showing marked substrate activation. With NAD+ as the variable substrate, double-reciprocal plots were linear, and apparently parallel. Double-reciprocal plots with enzyme modified with disulfiram (tetraethylthiuram disulphide) or iodoacetamide, such that at pH 8.0 the activity was decreased to 50% of the control value, showed no substrate activation, and the plots were linear. At pH 7.0, the kinetic parameters Vmax. and Km NAD+- for the oxidation of acetaldehyde and butyraldehyde by the native enzyme are almost identical. Formaldehyde and propionaldehyde show the same apparent maximum rate. Aldehyde dehydrogenase is able to catalyse the hydrolysis of p-nitrophenyl esters. This esterase activity was stimulated by both NAD+ and NADH, the maximum rate for the NAD+ stimulated esterase reaction being roughly equal to the maximum rate for the oxidation of aldehydes. The mechanistic implications of the above behaviour are discussed.
Article
Sheep liver cytoplasmic aldehyde dehydrogenase was purified to homogeneity to give a sample with a specific activity of 380 nmol NADH min(-1) mg(-1). An amino acid analysis of the enzyme gave results similar to those reported for aldehyde dehydrogenases from other sources. The isoelectric point was at pH 5.25 and the enzyme contained no significant amounts of metal ions. On the binding of NADH to the enzyme there is a shift in absorption maximum of NADH to 344 nm, and a 5.6-fold enhancement of nucleotide fluorescence. The protein fluorescence (lambdaexcit = 290 nm, lambdaemisson = 340 nm) is quenched on the binding of NAD+ and NADH. The enhancement of nucleotide fluorescence on the binding of NADH has been utilised to determine the dissociation constant for the enzyme . NADH complex (Kd = 1.2 +/- 0.2 muM). A Hill plot of the data gave a straight line with a slope of 1.0 +/- 0.3 indicating the absence of co-operative effects. Ellman's reagent reacted only slowly with the enzyme but in the presence of sodium dodecylsulphate complete reaction occurred within a few minutes to an extent corresponding to 36 thiol groups/enzyme. Molecular weights were determined for both cytoplasmic and mitochondrial aldehyde dehydrogenases and were 212 000 +/- 8 000 and 205 000 respectively. Each enzyme consisted of four subunits with molecular weight of 53 000 +/- 2 000. Properties of the cytoplasmic and mitochondrial aldehyde dehydrogenases from sheep liver were compared with other mammalian liver aldehyde dehydrogenases.
Article
In this review, an attempt is made to present all the relevant information in the scientific literature concerning the pharmacogenesis of the disulfiram-ethanol reaction. The following conclusions can reasonably be drawn. (1) Both disulfiram and its reduced derivative, diethyldithiocarbamate (DDC), are important to an understanding of the DER. Processes exist for the interconversion of these two compounds in vivo. (2) The prime cause of the DER is inactivation of aldehyde dehydrogenase by the sulfhydryl reagent, disulfiram, and the consequent accumulation of toxic levels of acetaldehyde following the ingestion of alcohol. (3) One effect of acetaldehyde is the release of norepinephrine, but, paradoxically, the DER is characterized by hypotension. This is thought to be due to blockage of the synthesis of norepinephrine at the dopamine β hydroxylase stage by DDC. (4) Any success achieved by disulfiram therapy (particularly when involving implantation) is probably due more to psychological factors than to the DER itself.
Article
Article
Preparations of sheep liver cytoplasmic aldehyde dehydrogenase obtained by published methods were found by analytical isoelectric focusing in the pH range 5--8 to contain 5--10% by weight of the mitochondrial aldehyde dehydrogenase. Under the conditions used the pI of the cytoplasmic enzyme is 6.2 and that of the mitochondrial enzyme 6.6. The mitochondrial enzyme can be removed from the preparation by selective precipitation of the cytoplasmic enzyme with (NH4)2SO4. Kinetic experiments and inhibition experiments with disulfiram show that the properties of the two sheep liver enzymes are so different that the presence of 10% mitochondrial enzyme in preparations of the cytoplasmic enzyme can introduce serious errors into results. Our results suggest that the presence of 10 microM-disulfiram in assays may completely inactivate the pure cytoplasmic enzyme. This result is in contrast with a previous report [kitson (1978) Biochem. U. 175, 83--90].
Article
The effect of disulfiram, [1-14C]disulfiram and some other thiol reagents on the activity of cytoplasmic aldehyde dehydrogenase from sheep liver was studied. The results are consistent with a rapid covalent interaction between disulfiram and the enzyme, and inconsistent with the notion that disulfiram is a reversible competitive inhibitor of cytoplasmic aldehyde dehydrogenase. There is a non-linear relationship between loss of about 90% of the enzyme activity and amount of disulfiram added; possible reasons for this are discussed. The remaining approx. 10% of activity is relatively insensitive to disulfiram. It is found that modification of only a small number of groups (one to two) per tetrameric enzyme molecule is responsible for the observed loss of activity. The dehydrogenase activity of the enzyme is affected more severely by disulfiram than is the esterase activity. Negatively charged thiol reagents have little or no effect on cytoplasmic aldehyde dehydrogenase. 2,2'-Dithiodipyridine is an activator of the enzyme.
Article
1. The activity of bovine liver glutamate dehydrogenase incubated with pyridoxal 5'-phosphate declined to a steady value reached within 30--60 min. The residual activity depended on the concentration of modifier up to about 5 mM. Above this concentration, however, no further inactivation was produced. The minimum activity obtainable in such incubations was 6--7% of the initial value. 2. Km values of the modified enzyme were unaltered, whereas Vmax. was decreased. 3. Activity was fully regained on dialysis against 0.1 M-potassium phosphate buffer. 4. Reduction with borohydride rendered the inactivation permanent but did not alter its extent. 5. Enzyme permanently inactivated in this way to the extent of 90% and dialysed was re-treated with pyridoxal 5'-phosphate. In this second cycle activity declined from 10 to 1% of the original activity. 6. This strongly suggests that the failure to achieve complete inactivation in a single cycle reflects a reversible equilibrium between inactive Schiff base, i.e. covalently modified enzyme, and a non-covalent complex. 7. The re-inactivation reaction occurring on dilution was demonstrated directly and a first-order rate constant obtained (0.048 min-1). This, in conjunction with an estimate of the forward rate constant for Schiff-base formation, obtained by approximate pseudo-first-order analysis of inactivation at varied modifier concentrations, gives a predicted minimum activity very close to that actually obtained in a single cycle of treatment. 8. The dissociation constant of the non-covalent complex is given by two methods as 0.90 and 1.59mM. 9. The results indicate that covalent modification with pyridoxal 5'-phosphate completely abolishes the activity of glutamate dehydrogenase.
Article
1.1. The distribution of aldehyde dehydrogenases in sheep liver was studied. Activity was found in the cytoplasm, mitochondria and microsomes.2.2. Cross-contamination of activities from different subcellular fractions, during the isolation procedures used, was shown to be insignificant. Accordingly, the level of aldehyde dehydrogenase activity found in each fraction should reflect the distribution pattern in vivo.3.3. Aldehyde dehydrogenases from the cytoplasm and mitochondria were isolated and some of their catalytic properties examined. The results show that the enzymes from the two fractions are not identical.