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 15 毫秒
1.
The interactions of rabbit muscle glyceraldehyde-3-phosphate dehydrogenase with NAD+ and with its fluorescent derivative 1, N6-etheno-adenine dinucleotide were investigated using a variety of spectroscopic methods. These techniques included: difference spectroscopy, circular dichroism, fluorescence and circular polarized luminescence. It was found that the greatest structural change in the protein tetramer occurs upon binding of the first mole of coenzyme. We have also demonstrated that progressive structural changes occur at the adenine subsite in the NAD+ binding site as a function of coenzyme saturation. These conformational changes are probably responsible for the progressive decrease in the affinity towards the coenzyme. It was also found that every NAD+ molecule induces the same conformational change of the nicotinamide subsite. These results offer a molecular explanation for the negative co-operativity in the binding of the coenzyme, without a change in the catalytic power of the NAD+ site as a function of coenzyme saturation. These results also offer a new explanation for the fact that enzyme exhibits half-of-the-sites reactivity towards certain ligands and full-site reactivity towards others. It is suggested that those ligands interacting at the adenine subsite of the NAD+ binding site induce the half-of-the-sites reactivity.Our results support the view that both the negative co-operativity in coenzyme binding and half-of-the-sites reactivity are due to ligand-induced conformational changes on an a priori symmetric glyceraldehyde-3-phosphate dehydrogenase molecule.  相似文献   

2.
The interaction of 3-aminopyridine-adenine dinucleotide, an NAD + 2 analogue which is fluorescent at the pyridine end of the molecule, with rabbit muscle glyceraldehyde-3-phosphate dehydrogenase was investigated. The fluorescence properties of the AAD+ molecule were used to monitor the nicotinamide subsites ou the GPDHase tetramer, the fluorescent aminopyridine moiety of the molecule serving as an intrinsic probe. Although the binding of AAD+ wag found to be negatively co-operative, no conformational changes induced at the nicotinamide subsite upon coenzyme binding were found to be transmitted to neighboring subunits. These findings, in conjunction with our earlier findings and with the observation that different NAD+ analogues which differ in the chemistry of the pyridine moiety bind with different extents of co-operativity, enable us to offer specific roles for the nicotinamide and the adenine subsites in generating the negative co-operativity.It is suggested that the structure of the pyridine moiety of the coenzyme controls the mode of binding of the pyridine moiety to the nicotinamide subsite. This, in turn, controls the orientation of the adenine moiety with respect to its subsite, thereby determining the mode of the interactions between the adenine and its binding domain. As the propagation of conformational changes caused by these interactions to neighboring subunits is believed to be the cause of the negative co-operativity exhibited by this enzyme towards coenzyme binding, the structure of the pyridine moiety controls this phenomenon.  相似文献   

3.
Acetone was found to form a dead-end ternary complex with horse liver alcohol dehydrogenase and oxidized nicotinamide adenine dinucleotide (NAD+) when the reactants were incubated for a long time at relatively high concentrations. The complex formation was demonstrated by measuring the increase in absorbance at 320 nm, the quenching of protein fluorescence, and the loss of enzyme activity. Since acetone is a substrate of liver alcohol dehydrogenase, and the presence of acetaldehyde or pyrazole prevents acetone from forming the dead-end complex with liver alcohol dehydrogenase and NAD+, the acetone molecule in the complex may be bound to the substrate binding site of liver alcohol dehydrogenase. The dissociation of the complex was demonstrated by prolonged dialysis or by addition of reduced nicotinamide adenine dinucleotide (NADH) and iso-butyramide. A modified nicotinamide adenine dinucleotide was obtained as a main product from the dead-end complex after dissociation of the complex or denaturation of the apoenzyme. The modified nicotinamide adenine dinucleotide was found to exhibit an absorption spectrum similar to that of NADH; however, it was not oxidizable by liver alcohol dehydrogenase in the presence of acetaldehyde and exhibited no fluorescence.  相似文献   

4.
As shown by X-ray crystallography, horse liver alcohol dehydrogenase undergoes a global conformational change upon binding of NAD+ or NADH, involving a rotation of the catalytic domain relative to the coenzyme binding domain and the closing up of the active site to produce a catalytically efficient enzyme. The conformational change requires a complete coenzyme and is affected by various chemical or mutational substitutions that can increase the catalytic turnover by altering the kinetics of the isomerization and rate of dissociation of coenzymes. The binding of NAD+ is kinetically limited by a unimolecular isomerization (corresponding to the conformational change) that is controlled by deprotonation of the catalytic zinc-water to produce a negatively-charged zinc-hydroxide, which can attract the positively-charged nicotinamide ring. The deprotonation is facilitated by His-51 acting through a hydrogen-bonded network to relay the proton to solvent. Binding of NADH also involves a conformational change, but the rate is very fast. After the enzyme binds NAD+ and closes up, the substrate displaces the hydroxide bound to the catalytic zinc; this exchange may involve a double displacement reaction where the carboxylate group of a glutamate residue first displaces the hydroxide (inverting the tetrahedral coordination of the zinc), and then the exogenous ligand displaces the glutamate. The resulting enzyme-NAD+-alcoholate complex is poised for hydrogen transfer, and small conformational fluctuations may bring the reactants together so that the hydride ion is transferred by quantum mechanical tunneling. In the process, the nicotinamide ring may become puckered, as seen in structures of complexes of the enzyme with NADH. The conformational changes of alcohol dehydrogenase demonstrate the importance of protein dynamics in catalysis.  相似文献   

5.
Glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides utilizes either NAD+ or NADP+ as coenzyme. Kinetic studies showed that NAD+ and NADP+ interact with different enzyme forms (Olive, C., Geroch, M. E., and Levy, H. R. (1971) J. Biol. Chem.246, 2047–2057). In the present study the techniques of fluorescence quenching and fluorescence enhancement were used to investigate the interaction between Leuconostoc mesenteroides glucose-6-phosphate dehydrogenase and coenzymes. In addition, kinetic studies were performed to examine interaction between the enzyme and various coenzyme analogs. The maximum quenching of protein fluorescence is 5% for NADP+ and 50% for NAD+. The dissociation constant for NADP+, determined from fluorescence quenching measurements, is 3 μm, which is similar to the previously determined Km of 5.7 μm and Ki of 5 μm. The dissociation constant for NAD+ is 2.5 mm, which is 24 times larger than the previously determined Km of 0.106 mm. Glucose 1-phosphate, a substrate-competitive inhibitor, lowers the dissociation constant and maximum fluorescence quenching for NAD+ but not for NADP+. This suggests that glucose 6-phosphate may act similarly and thus play a role in enabling the enzyme to utilize NAD+ under physiological conditions. When NADPH binds to the enzyme its fluorescence is enhanced 2.3-fold. The enzyme was titrated with NADPH in the absence and presence of NAD+; binding of these two coenzymes is competitive. The dissociation constant for NADPH from these measurements is 24 μm; the previously determined Ki is 37.6 μm. The dissociation constant for NAD′ is 2.8 mm, in satisfactory agreement with the value obtained from protein fluorescence quenching measurements. Various compounds which resemble either the adenosine or the nicotinamide portion of the coenzyme structure are coenzyme-competitive inhibitors; 2′,5′-ADP, the most inhibitory analog tested, gives NADP+-competitive and NAD+-noncompetitive inhibition, consistent with the kinetic mechanism previously proposed. By using pairs of coenzyme-competitive inhibitors it was shown in kinetic studies that the two portions of the NAD+ structure cannot be accommodated on the enzyme simultaneously unies they are covalently linked. Fluorescence studies showed that there are both “buried” and “exposed” tryptophan residues in the enzyme structure.  相似文献   

6.
Riboswitches are important gene regulatory elements frequently encountered in bacterial mRNAs. The recently discovered nadA riboswitch contains two similar, tandemly arrayed aptamer domains, with the first domain possessing high affinity for nicotinamide adenine dinucleotide (NAD+). The second domain which comprises the ribosomal binding site in a putative regulatory helix, however, has withdrawn from detection of ligand-induced structural modulation thus far, and therefore, the identity of the cognate ligand and the regulation mechanism have remained unclear. Here, we report crystal structures of both riboswitch domains, each bound to NAD+. Furthermore, we demonstrate that ligand binding to domain 2 requires significantly higher concentrations of NAD+ (or ADP retaining analogs) compared to domain 1. Using a fluorescence spectroscopic approach, we further shed light on the structural features which are responsible for the different ligand affinities, and describe the Mg2+-dependent, distinct folding and pre-organization of their binding pockets. Finally, we speculate about possible scenarios for nadA RNA gene regulation as a putative two-concentration sensor module for a time-controlled signal that is primed and stalled by the gene regulation machinery at low ligand concentrations (domain 1), and finally triggers repression of translation as soon as high ligand concentrations are reached in the cell (domain 2).  相似文献   

7.
The ethanologenic bacterium Zymomonas mobilis ZM4 is of special interest because it has a high ethanol yield. This is made possible by the two alcohol dehydrogenases (ADHs) present in Z. mobilis ZM4 (zmADHs), which shift the equilibrium of the reaction toward the synthesis of ethanol. They are metal-dependent enzymes: zinc for zmADH1 and iron for zmADH2. However, zmADH2 is inactivated by oxygen, thus implicating zmADH2 as the component of the cytosolic respiratory system in Z. mobilis. Here, we show crystal structures of zmADH2 in the form of an apo-enzyme and an NAD+-cofactor complex. The overall folding of the monomeric structure is very similar to those of other functionally related ADHs with structural variations around the probable substrate and NAD+ cofactor binding region. A dimeric structure is formed by the limited interactions between the two subunits with the bound NAD+ at the cleft formed along the domain interface. The catalytic iron ion binds near to the nicotinamide ring of NAD+, which is likely to restrict and locate the ethanol to the active site together with the oxidized Cys residue and several nonpolar bulky residues. The structures of the zmADH2 from the proficient ethanologenic bacterium Z. mobilis, with and without NAD+ cofactor, and modeling ethanol in the active site imply that there is a typical metal-dependent catalytic mechanism.  相似文献   

8.
The crystal structure of the NADH:quinone oxidoreductase PA1024 has been solved in complex with NAD+ to 2.2 Å resolution. The nicotinamide C4 is 3.6 Å from the FMN N5 atom, with a suitable orientation for facile hydride transfer. NAD+ binds in a folded conformation at the interface of the TIM‐barrel domain and the extended domain of the enzyme. Comparison of the enzyme‐NAD+ structure with that of the ligand‐free enzyme revealed a different conformation of a short loop (75–86) that is part of the NAD+‐binding pocket. P78, P82, and P84 provide internal rigidity to the loop, whereas Q80 serves as an active site latch that secures the NAD+ within the binding pocket. An interrupted helix consisting of two α‐helices connected by a small three‐residue loop binds the pyrophosphate moiety of NAD+. The adenine moiety of NAD+ appears to π–π stack with Y261. Steric constraints between the adenosine ribose of NAD+, P78, and Q80, control the strict specificity of the enzyme for NADH. Charged residues do not play a role in the specificity of PA1024 for the NADH substrate.  相似文献   

9.
J M Thornton  P M Bayley 《Biopolymers》1977,16(9):1971-1986
A study of the conformational states of the dinucleotide coenzyme NAD+ has been made using semiempirical energy calculations. Taking low-energy mononucleotide structures as starting conformations, energy minimizations have been performed. The lowest energy states are stacked structures, with interactions between the adenine and nicotinamide rings. Some structures show stabilization gained from electrostatic attractions between the positively charged nicotinamide and negatively charged phosphate oxygens. These predictions correlate well with the available experimental data.  相似文献   

10.
Butanedione in the presence of borate buffer reversibly inhibits Rhodospirillum rubrum chromatophore transhydrogenase complex and the separated membrane-bound and soluble factor components of the complex. NADP+ completely protected against inactivation of the membrane-bound component, whereas NAD+ was without effect. Soluble factor was maximally protected only partially by either NAD+ or NADP+, but a mixture of the substrates afforded complete protection. NADP+-dependent association of soluble factor with factor-depleted membranes was markedly decreased after incubation of membranes with butanedione in the absence, but not in the presence, of NADP+. Soluble factor was bound to agarose-NAD and was eluted by NAD+, but not by NADP+. These results demonstrate the presence of at least three nicotinamide adenine dinucleotide binding sites on R. rubrum transhydrogenase complex, including separate NADP and NAD binding sites on soluble factor and a NADP binding site on the membrane-bound component.  相似文献   

11.
The epimerase MoeE5 from Streptomyces viridosporus converts UDP-glucuronic acid (UDP-GlcA) to UDP-galacturonic acid (UDP-GalA) to provide the first sugar in synthesizing moenomycin, a potent inhibitor against bacterial peptidoglycan glycosyltransferases. The enzyme belongs to the UDP-hexose 4-epimerase family, and uses NAD+ as its cofactor. Here we present the complex crystal structures of MoeE5/NAD+/UDP-GlcA and MoeE5/NAD+/UDP-glucose, determined at 1.48 Å and 1.66 Å resolution. The cofactor NAD+ is bound to the N-terminal Rossmann-fold domain and the substrate is bound to the smaller C-terminal domain. In both crystals the C4 atom of the sugar moiety of the substrate is in close proximity to the C4 atom of the nicotinamide of NAD+, and the O4 atom of the sugar is also hydrogen bonded to the side chain of Tyr154, suggesting a productive binding mode. As the first complex structure of this protein family with a bound UDP-GlcA in the active site, it shows an extensive hydrogen-bond network between the enzyme and the substrate. We further built a model with the product UDP-GalA, and found that the unique Arg192 of MoeE5 might play an important role in the catalytic pathway. Consequently, MoeE5 is likely a specific epimerase for UDP-GlcA to UDP-GalA conversion, rather than a promiscuous enzyme as some other family members.  相似文献   

12.
Rape alcohol dehydrogenase is competitively inhibited with respect to NAD by nicotinamide, as well as by compounds containing adenine (adenine, adenosine, AMP, ADP, ATP). Adenine and adenosine are bound more firmly to the enzyme than nicotinamide. The two types of compound, as component parts of the NAD coenzyme, are bound to different sites on the enzyme. Adenine and adenosine compete for the adenine nucleotide bonding site, but they do not compete for the o-phenanthroline bonding site. Nicotinamide competes with o-phenanthroline for the binding site at which the metal is apparently present.  相似文献   

13.
Betaine aldehyde dehydrogenase (BADH EC 1.2.1.8) catalyzes the irreversible oxidation of betaine aldehyde to glycine betaine using NAD+ as a coenzyme. Porcine kidney BADH (pkBADH) follows a bi‐bi ordered mechanism in which NAD+ binds to the enzyme before the aldehyde. Previous studies showed that NAD+ induces complex and unusual conformational changes on pkBADH and that potassium is required to maintain its quaternary structure. The aim of this work was to analyze the structural changes in pkBADH caused by NAD+ binding and the role played by potassium in those changes. The pkBADH cDNA was cloned and overexpressed in Escherichia coli, and the protein was purified by affinity chromatography using a chitin matrix. The pkBADH/NAD+ interaction was analyzed by circular dichroism (CD) and by isothermal titration calorimetry (ITC) by titrating the enzyme with NAD+. The cDNA has an open reading frame of 1485 bp and encodes a protein of 494 amino acids, with a predicted molecular mass of 53.9 kDa. CD data showed that the binding of NAD+ to the enzyme caused changes in its secondary structure, whereas the presence of K+ helps maintain its α‐helix content. K+ increased the thermal stability of the pkBADH‐NAD+ complex by 5.3°C. ITC data showed that NAD+ binding occurs with different association constants for each active site between 37.5 and 8.6 μM. All the results support previous data in which the enzyme incubation with NAD+ provoked changes in reactivity, which is an indication of slow conformational rearrangements of the active site.  相似文献   

14.
A range of biomimetic analogues of the nicotinamide nucleotide coenzymes NAD(P)(H) have been developed based on the structure of a triazine dye template. These biomimetic redox coenzymes are relatively straightforward and inexpensive to synthesise and display NAD+-like activity with different dehydrogenases, despite their apparently minimal structural similarity to the native coenzyme NAD+. Horse liver alcohol dehydrogenase oxidises butan-1-ol, using the most active biomimetic coenzyme (Nap 1), with a k cat value an order of magnitude lower and a K m for the coenzyme two orders of magnitude higher than those using native NAD+. The enzymatically reduced biomimetic coenzymes may be reoxidised by phenazine methosulfate. We believe that these coenzymes may find applications in biotransformations and biosensors, and in the development of biomimetic catalysts where the redox enzyme itself is replaced by a synthetic binding site. Received: 26 October 1998 / Received revision: 25 January 1999 / Accepted: 31 January 1999  相似文献   

15.
Liver alcohol dehydrogenase (LADH; E.C. 1.1.1.1) provides an excellent system for probing the role of binding interactions with NAD+ and alcohols as well as with NADH and the corresponding aldehydes. The enzyme catalyzes the transfer of hydride ion from an alcohol substrate to the NAD+ cofactor, yielding the corresponding aldehyde and the reduced cofactor, NADH. The enzyme is also an excellent catalyst for the reverse reaction. X-ray crystallography has shown that the NAD+ binds in an extended conformation with a distance of 15 Å between the buried reacting carbon of the nicotinamide ring and the adenine ring near the surface of the horse liver enzyme. A major criticism of X-ray crystallographic studies of enzymes is that they do not provide dynamic information. Such data provide time-averaged and space-averaged models. Significantly, entries in the protein data bank contain both coordinates as well as temperature factors. However, enzyme function involves both dynamics and motion. The motions can be as large as a domain closure such as observed with liver alcohol dehydrogenase or as small as the vibrations of certain atoms in the active site where reactions take place. Ternary complexes produced during the reaction of the enzyme binary entity, E-NAD+, with retinol (vitamin A alcohol) lead to retinal (vitamin A aldehyde) release and the enzyme binary entity E-NADH. Retinal is further metabolized via the E-NAD+-retinal ternary complex to retinoic acid (vitamin A acid). To unravel the mechanistic aspects of these transformations, the kinetics and energetics of interconversion between various ternary complexes are characterized. Proton transfers along hydrogen bond bridges and NADH hydride transfers along hydrophobic entities are considered in some detail. Secondary kinetic isotope effects with retinol are not particularly large with the wild-type form of alcohol dehydrogenase from horse liver. We analyze alcohol dehydrogenase catalysis through a re-examination of the reaction coordinates. The ground states of the binary and ternary complexes are shown to be related to the corresponding transition states through topology and free energy acting along the reaction path.  相似文献   

16.
17.
The pH dependence of the 13C chemical shifts for nicotinamide adenine dinucleotide (NAD+), thionicotinamide adenine dinucleotide (TNAD+), pyridine adenine dinucleotide (PyrAD+), N-methyl-nicotinamide adenine dinucleotide (N-Me-NAD+), acetylpyridine adenine dinucleotide (AcPyAD+), nicotinamide hypoxanthine dinucleotide (NHD+), and nicotinamide adenine dinucleotide phosphate (NADP+) are reported. In these analogs the 13C chemical shifts of the pyridinium moiety reflect the pKa of the opposing purine base, while the 13C chemical shift dependence on pD for the pyridinium carbons of nicotinamide mononucleotide (NMN+) and adenosine monophosphate (AMP), 1,4-dihydronicotinamide adenine dinucleotide (NADH), 1,4-dihydronicotinamide adenine dinucleotide phosphate (NADPH), and nicotinic acid adenine dinucleotide (N(a)AD+) are not influenced by the adenine ring in the pD range tested. Through the use of 13C-labeled NAD+, the source of the pH dependence of the 13C chemical shifts was shown to be intramolecular in origin. However, serious doubt is cast on the utility of employing the pD dependence of chemical shift data to determine the nature of solution conformers or their relative populations.  相似文献   

18.
Ferredoxin-NADP+ reductase (FNR) catalyses the electron transfer from ferredoxin to NADP+ via its flavin FAD cofactor. A molecular dynamics theoretical approach is applied here to visualise the transient catalytically competent interaction of Anabaena FNR with its coenzyme, NADP+. The particular role of some of the residues identified as key in binding and accommodating the 2′P-AMP moiety of the coenzyme is confirmed in molecular terms. Simulations also indicate that the architecture of the active site precisely contributes to the orientation of the N5 of the FAD isoalloxazine ring and the C4 of the coenzyme nicotinamide ring in the conformation of the catalytically competent hydride transfer complex and, therefore, contributes to the efficiency of the process. In particular, the side chain of the C-terminal Y303 in Anabaena FNR appears key to providing the optimum geometry by reducing the stacking probability between the isoalloxazine and nicotinamide rings, thus providing the required co-linearity and distance among the N5 of the flavin cofactor, the C4 of the coenzyme nicotinamide and the hydride that has to be transferred between them. All these factors are highly related to the reaction efficiency, mechanism and reversibility of the process.  相似文献   

19.
Aminoaldehyde dehydrogenases (AMADHs, EC 1.2.1.19) belong to the large aldehyde dehydrogenase (ALDH) superfamily, namely, the ALDH9 family. They oxidize polyamine-derived ω-aminoaldehydes to the corresponding ω-amino acids. Here, we report the first X-ray structures of plant AMADHs: two isoenzymes, PsAMADH1 and PsAMADH2, from Pisum sativum in complex with β-nicotinamide adenine dinucleotide (NAD+) at 2.4 and 2.15 Å resolution, respectively. Both recombinant proteins are dimeric and, similarly to other ALDHs, each monomer is composed of an oligomerization domain, a coenzyme binding domain and a catalytic domain. Each subunit binds NAD+ as a coenzyme, contains a solvent-accessible C-terminal peroxisomal targeting signal (type 1) and a cation bound in the cavity close to the NAD+ binding site. While the NAD+ binding mode is classical for PsAMADH2, that for PsAMADH1 is unusual among ALDHs. A glycerol molecule occupies the substrate binding site and mimics a bound substrate. Structural analysis and substrate specificity study of both isoenzymes in combination with data published previously on other ALDH9 family members show that the established categorization of such enzymes into distinct groups based on substrate specificity is no more appropriate, because many of them seem capable of oxidizing a large spectrum of aminoaldehyde substrates. PsAMADH1 and PsAMADH2 can oxidize N,N,N-trimethyl-4-aminobutyraldehyde into γ-butyrobetaine, which is the carnitine precursor in animal cells. This activity highly suggests that in addition to their contribution to the formation of compatible osmolytes such as glycine betaine, β-alanine betaine and γ-aminobutyric acid, AMADHs might participate in carnitine biosynthesis in plants.  相似文献   

20.
Comparison of the calculated structures for the transition states of the two reactions catalysed by chorismate mutase prephenate dehydrogenase suggests that both reactions could be catalysed at a common active site. Kinetic data for the enzyme from Aerobacteraerogenes are consistent with this possibility. On the basis of these theoretical and experimental data a model for a common active site is developed. In the model, the transition state for each reaction is bound to the enzyme via both of the two substrate carboxyl groups, and can also interact with the coenzyme nicotinamide adenine dinucleotide through a hydrogen bond between the amide moiety of the nicotinamide ring and the hydroxyl group of the substrate. Chorismate, prephenate and 4-hydroxyphenylpyruvate in their ground states form the same hydrogen bond to the coenzyme, but are bound to the enzyme via a single carboxyl group only. The additional bond formed between the enzyme and the transition state structures thus provides the transition state stabilization required for catalysis of both reactions.  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