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 78 毫秒
1.
The catalytic site ofEscherichia coli F1-ATPase is reviewed in terms of structure and function. Structural prediction, biochemical analyses, and mutagenesis experiments suggest that the catalytic site is formed primarily by residues 137–335 of -subunit. Subdomains of the site involved in phosphate-bond cleavage/synthesis and adenine-ring binding are discussed. Ambiguities inherent in steady-state catalytic measurements due to catalytic site cooperativity are discussed, and the advantages of pre-steady-state (unisite) techniques are emphasized. The emergence of a single high-affinity catalytic site occurs as a result of F1-oligomer assembly. Measurements of unisite catalysis rate and equilibrium constants, and their modulation by varied pH, dimethylsulfoxide, and mutations, are described and conclusions regarding the nature of the high-affinity catalytic site and mechanism of catalysis are presented.  相似文献   

2.
3.
The F1 part of the ATP synthase contains 6 nucleotide binding sites, four of which can be occupied and covalently labeled with 8-azido-adenine nucleotides. The other two sites contain tightly bound nucleotides that cannot be replaced by 8-azido-adenine nucleotides. Of the four exchangeable sites two are directly ivolved in catalysis and these are located on -subunits, while the other two are located at - interfaces and have probably a regulatory role by influencing the affinity of the catalytic sites for substrate and product. When only one catalytic site contains substrate the affinity is very high, the rate of hydrolysis is slow, and the dissociation of products is even slower (single-site catalysis). When the second site also becomes occupied, the affinity decreases enormously, and the rate of hydrolysis and dissociation of products increases several orders of magnitude. When, however, the second site is occupied by substrate in such a way that turnover is not possible at this site (e.g., covalent linkage of nitreno-ATP), the first site is no longer active, apart from the very slow single-site catalysis. The two nonexchangeable, tightly bound nucleotides that cannot be replaced by 8-azido-nucleotides, can be replaced by 2-azido-nucleotides, due to their anticonfiguration. This anticonfiguration of the substrate is also required for binding with high affinity to a catalytic site. A picture emerges in which one of the three - pairs of F1 contains tightly bound, nonexchangeable nucleotides, while the other two contain both one catalytic site (on ) and one regulatory site (at the - interface). Cooperativity exists both between the two catalytic sites and between the catalytic and the regulatory sites.  相似文献   

4.
The mitochondrial ATP synthases shares many structural and kinetic properties with bacterial and chloroplast ATP synthases. These enzymes transduce the energy contained in the membrane's electrochemical proton gradients into the energy required for synthesis of high-energy phosphate bonds. The unusual three-fold symmetry of the hydrophilic domain, F1, of all these synthases is striking. Each F1 has three identical subunits and three identical subunits as well as three additional subunits present as single copies. The catalytic site for synthesis is undoubtedly contained in the subunit or an , interface, and thus each enzyme appears to contain three identical catalytic sites. This review summarizes recent isotopic and kinetic evidence in favour of the concept, originally proposed by Boyer and coworkers, that energy from the proton gradient is exerted not directly for the reaction at the catalytic site, but rather to release product from a single catalytic site. A modification of this binding change hypotheses is favored by recent data which suggest that the binding change is due to a positional change in all three subunits relative to the remaining subunits of F1 and F0 and that the vector of rotation is influenced by energy. The positional change, or rotation, appears to be the slow step in the process of catalysis and it is accelerated in all F1F0 ATPases studied by substrate binding and by the proton gradient. However, in the mammalian mitochondrial enzyme, other types of allosteric rate regulation not yet fully elucidated seem important as well.  相似文献   

5.
The thermal denaturation of the simple, redox-active iron protein rubredoxin is characterized by a slow, irreversible decay of the characteristic red color of the iron center at elevated temperatures in the presence of oxygen at pH 7.8. The denaturation rate is essentially constant and the time period for complete bleaching is nearly independent of protein concentration. These two characteristics of the kinetics can be fit by a simple self-catalyzed kinetics model consisting of the combination of a first-order decay and catalysis by some product of that decay, i.e., dP/dt=k 1[A]+(k 2[P][A])/(K m+[A]), where A is native rubredoxin, P, is unspecified product, k 1 is a first-order rate constant, and k 2 and K m are the catalytic constants. In order for the second term to be of this simple form over the full course of a decay, the model must include the condition that the reaction is effectively irreversible. This model has properties which suggest other biological roles in regulation (changes in k 1 or k 2 can dramatically modulate the kinetics), in timing (titer-independent fixed reaction time), and in self-activation reactions. At one extreme (k1 k2) the kinetics becomes exponential, but at the other extreme (k2 k1) they show a dramatic and rapid terminal increase after a lag period. Some obvious possible roles in the kinetics of programmed cell death, prion disease, and protease autoactivation are discussed.  相似文献   

6.
7.
The catalytic portion (F1) of ATP synthases have the subunit composition 3, 3, , , . This composition imparts structural asymmetry to the entire complex that results in differences in nucleotide binding affinity among the six binding sites. Evidence that two or more sites participate in catalysis, alternating their properties, led to the notion that the interactions of individual pairs with the small subunits must change as binding site properties alternate. A rotation of the subunit within the 33 hexamer has been proposed as a means of alternating the properties of catalytic sites. Evidence argues that the rotation of the complete subunit during ATP hydrolysis is not mandatory for activity. The subunit of chloroplast F1 may be cleaved into three large fragments that remain bound to F1. This cleavage enhances ATPase activity without loss of evidence of site-site interactions. Complexes of 33 have been shown to have significant ATPase activity in the absence of . Mg2+ATP affects the interaction of with the different subunits, and induces other changes in F1, but whether these changes are induced by catalysis, or are fast enough to be involved in the catalytic turnover of the enzyme has not been established. Likewise, changes in structure and in binding site properties induced in thylakoid membrane bound CF1 by formation of an electrochemical proton gradient may activate the enzyme rather than be apart of catalysis. Mechanisms other than rotary catalysis should be considered.  相似文献   

8.
The approach of CO2 to a series of active site model complexes of human carbonic anhydrase II (HCAII) and its catalytic hydration to bicarbonate anion have been investigated using semiempirical MO theory (AM1). The results show that direct nucleophilic attack of zinc-bound hydroxide to the substrate carbon occurs in each model system. Further rearrangement of the bicarbonate complex thus formed via a rotation-like movement of the bicarbonate ligand can only be found in active site model systems that include at least one additional water molecule. Further refinement of the model complex by adding a methanol molecule to mimic Thr-199 makes this process almost activationless. The formation of the final bicarbonate complex by an internal (intramolecular) proton transfer is only possible in the simplest of all model systems, namely {[Im3Zn(OH)]+·CO2}. The energy of activation for this process, however, is 36.8 kcal·mol–1 and thus too high for enzymatic catalysis. Therefore, we conclude that within the limitations of the model systems presented and the level of theory employed, the overall mechanism for the formation of the bicarbonate complex comprises an initial direct nucleophilic attack of zinc-bound hydroxide to carbon dioxide followed by a rotation-like rearrangement of the bicarbonate ligand via a penta-coordinate Zn2+ transition state structure, including the participation of an extra active site water molecule.Electronic Supplementary Material available.  相似文献   

9.
The main parts of the central olfactory system are the bulb (OB), anterior nucleus (AON), and prepyriform cortex (PC). Each part consists of a mass of excitatory or inhibitory neurons that is modelled in its noninteractive state by a 2nd order ordinary differential equation (ODE) having a static nonlinearity. The model is called a KOe or a KOt set respectively; it is evaluated in the open loop state under deep anesthesia. Interactions in waking states are represented by coupled KO sets, respectivelyKI e (mutual excitation) andKI i (mutual inhibition). The coupledKI e andKI i sets form aKII set, which suffices to represent the dynamics of theOB, AON, andPC separately. The coupling of these three structures by both excitatory and inhibitory feedback loops forms aKIII set. The solutions to this high-dimensional system ofODEs suffice to simulate the chaotic patterns of the EEG, including the normal low-level background activity, the high-level relatively coherent bursts of oscillation that accompany reception of input to the bulb, and a degenerate state of an epileptic seizure determined by a toroidal chaotic attractor. An example is given of the Ruelle-Takens-Newhouse route to chaos in the olfactory system. Due to the simplicity and generality of the elements of the model and their interconnections, the model can serve as the starting point for other neural systems that generate deterministic chaotic activity.Supported by a grant MH06686 from the National Institute of Mental Health  相似文献   

10.
11.
Ultrasonic and calorimetric studies of small homogenously-sized DMPC unilamellar vesicles showed two thermal transitions at temperatures T c1 and T c2 (T c2 T c1 ); T c2 is close to the phase transition temperature, T c , of large vesicles. The process at T c2 is not a fusion of vesicles and is interpreted as characterizing an order-disorder transition essentially similar to that of large vesicles. The temperatures T c1 and T c2 become increasingly similar as the cholesterol content is increased, while the clusters at T c2 (85 lipid molecules in pure DMPC) increase in size up to approximately 180 lipid molecules at 12 mol% cholesterol. Incorporation of cholesterol thus brings about enhanced fluctuations in this model system of a membrane.Abbreviations DMPC dimyristoylphosphatidylcholine - SUV small unilamellar vesicles - LUV large unilamellar vesicles - MLV multilamellar vesicles  相似文献   

12.
The Graded Autocatalysis Replication Domain (GARD) model describes an origin of life scenario which involves non-covalent compositional assemblies, made of monomeric mutually catalytic molecules. GARD constitutes an alternative to informational biopolymers as a mechanism of primordial inheritance. In the present work, we examined the effect of mutations, one of the most fundamental mechanisms for evolution, in the context of the networks of mutual interaction within GARD prebiotic assemblies. We performed a systematic analysis analogous to single and double gene deletions within GARD. While most deletions have only a small effect on both growth rate and molecular composition of the assemblies, ~10% of the deletions caused lethality, or sometimes showed enhanced fitness. Analysis of 14 different network properties on 2,000 different GARD networks indicated that lethality usually takes place when the deleted node has a high molecular count, or when it is a catalyst for such node. A correlation was also found between lethality and node degree centrality, similar to what is seen in real biological networks. Addressing double knockout mutations, our results demonstrate the occurrence of both synthetic lethality and extragenic suppression within GARD networks, and convey an attempt to correlate synthetic lethality to network node-pair properties. The analyses presented help establish GARD as a workable alternative prebiotic scenario, suggesting that life may have begun with large molecular networks of low fidelity, that later underwent evolutionary compaction and fidelity augmentation.  相似文献   

13.
Self-reproduction and the ability to regulate their composition are two essential properties of terrestrial biotic systems. The identification of non-living systems that possess these properties can therefore contribute not only to our understanding of their functioning but also hint at possible prebiotic processes that led to the emergence of life. Growing lipid vesicles have been previously established as having the capacity to self-reproduce. Here it is demonstrated that vesicle self-reproduction can occur only at selected values of vesicle properties. We treat as an example a simple vesicle with membrane elastic properties defined by a membrane bending modulus and spontaneous curvature C0, whose volume variation depends on the membrane hydraulic permeability Lp and whose membrane area doubles in time Td. Vesicle self-reproduction is described as a process in which a growing vesicle first transforms its shape from a sphere into a budded shape of two spheres connected by a narrow neck, and then splits into two spherical daughter vesicles. We show that budded vesicle shapes can be reached only under the condition that TdLpC041.85. Thus, in a growing vesicle population containing vesicles of different composition, only the vesicles for which this condition is fulfilled can increase their number in a self-reproducing manner. The obtained results also suggest that at times much longer than Td the number of vesicles with their properties near the edge in the system parameter space defined by the minimum value of the product TdLpC04, will greatly exceed the number of any other vesicles.  相似文献   

14.
Summary This review concerns the catalytic sector of F1 factor of the H+-dependent ATPases in mitochondria (MF1), bacteria (BF1) and chloroplasts (CF1). The three types of Ft have many similarities with respect to the structural parameters, subunit composition and catalytic mechanism. An 33222 stoichiometry is now accepted for MF1 and BF1; the 22222 stoichiometry for CFI remains as matter of debate. The major subunits , and are equivalent in MF1, BF1 and CF1; this is not the case for the minor subunits and . The subunit of MFI corresponds to the subunit of BF1 and CF1, whereas the mitochondria) subunit equivalent to the subunit of BF1 and CF1 is probably the oligomycin sensitivity conferring protein (OSCP). The a assembly is endowed with ATPase activity, being considered as the catalytic subunit and y as a proton gate. On the other hand, the 6 and E subunits of BFI and CFI most probably act as links between the F1 and F0 sectors of the ATPase complex. The natural mitochondria) ATPase inhibitor, which is a separate protein loosely attached to MF1, could have its counterpart in the E subunit of BF1 and CF1. The generally accepted view that the catalytic subunit in the different F1 species is comes from a number of approaches, including chemical modification, specific photolabeling and, in the case of BF1, use of mutants. The a subunit also plays a central role in catalysis, since structural alteration of a by chemical modification or mutation results in loss of activity of the whole molecule of F1. The notion that the proton motive force generated by respiration is required for conformational changes of the F1 sector of the H+-ATPase complex has gained acceptance. During the course of ATP synthesis, conversion of bound ADP and Pi into bound ATP probably requires little energy input; only the release of the F1-bound ATP would consume energy. ADP and Pi most likely bind at one catalytic site of F1, while ATP is released at another site. This mechanism, which underlines the alternating cooperativity of subunits in F1, is supported by kinetic data and also by the demonstration of partial site reactivity in inactivation experiments performed with selective chemical modifiers. One obvious advantage of the alternating site mechanism is that the released ATP cannot bind to its original site. The chemistry of the condensation reaction of ADP and Pi to form ATP has not yet been elucidated. Although implicitly admitted, definite evidence that the condensation reaction does not involve a phosphorylated intermediate has been acquired recently by analysis of the stereochemical course of the phosphoric residue transfer in ATP synthesis or hydrolysis. Whereas the catalytic events of ATP synthesis are well understood, the regulatory mechanism, and particularly the role of the so-called inhibitory peptides, remain enigmatic.  相似文献   

15.
Summary In previous work on rat striated muscle cells a sliver-reducing component was found selectively localized at the terminal cistern/transverse tubule system (Tandler and Pellegrino de Iraldi 1989). To further investigate that problem we performed the Hg–Ag argentaffin reaction on a sarcoplasmic reticulum fraction from rat skeletal muscle. Circular profiles corresponding to vesicular structures were found outlined by silver grains. The number of silver stained vesicles were less than the total number vesicles stained by conventional procedures. The correlation between argentaffinities in the intact muscle fiber and their subcellular organelles indicated that the Hg–Ag reactive vesicles must be those derived from the terminal cisternae of the sarcoplasmic reticulum. The silver-reducing constituent aggregates in the presence of 1 mM CaCl2 or 0.5 M K cacodylate. The state of aggregation induced by Ca2+ was not affected by incubation with 0.5% Triton X-100 or by 2 mM EDTA, thus suggesting a localization at or near the membrane of the terminal cistern vesicle facing the junctional gap. In Laemmli SDS-acrylamide gels the Hg–Ag reaction stained all proteins in a manner similar to Coomasie blue. It is suggested that the selective histochemical staining is the result of differential reactivities due to steric requirements of the chemical reaction.  相似文献   

16.
Nitroaromatic compounds are important intermediates in organic synthesis. The classic method used to synthesize them is chemical nitration, which involves the use of nitric acid diluted in water or acetic acid, both harmful to the environment. With the development of green chemistry, environmental friendly enzyme catalysis is increasingly employed in chemical processes. In this work, we adopted a non-aqueous horseradish peroxidase (HRP)/NaNO2/H2O2 reaction system to study the structural characteristics of aromatic compounds potentially nitrated by enzyme catalysis, as well as the relationship between the charges on carbon atoms in benzene ring and the nitro product distribution. Investigation of various reaction parameters showed that mild reaction conditions (ambient temperature and neutral pH), plus appropriate use of H2O2 and NaNO2 could prevent inactivation of HRP and polymerization of the substrates. Compared to aqueous–organic co-solvent reaction media, the aqueous–organic two-liquid phase system had great advantages in increasing the dissolved concentration of substrate and alleviating substrate inhibition. Analysis of the aromatic compounds’ structural characteristics indicated that substrates containing substituents of NH2 or OH were readily catalyzed. Furthermore, analysis of the relationship between natural bond orbital (NBO) charges on carbon atoms in benzene ring, as calculated by the density functional method, and the nitro product distribution characteristics, demonstrated that the favored nitration sites were the ortho and para positions of substituents in benzene ring, similar to the selectivity of chemical nitration.  相似文献   

17.
Three characteristics have been demonstrated by the chemical analysis of bovine brain coated vesicles following removal of the coat proteins: a high protein content, a high cholesterol/lipid ratio and a high percentage of phosphatidylethanolamine amongst the phospholipids.The study of lipid bilayer organization and dynamics has been performed using the fluorescent probes pyrene and parinaric acid (cis and trans). This has allowed the study of both lateral mobility and rotational motion in the lipid bilayer of the coated and uncoated vesicles.Lateral mobility in the fluid phase of the lipid is slightly reduced by the presence of the clathrin coat, as indicated by the lower diffusion coefficient of pyrene in coated compared with uncoated vesicles.At all temperatures from 6° to 30°C, solid-phase domains, probed by trans parinaric acid, coexist with fluid-phase domains in the lipid bilayer. The temperature dependence of the parinaric acid lifetimes and of their amplitudes strongly suggests that the solid phase domains decrease in size with temperature, both in coated and uncoated vesicles.However, the difference in the value of the anisotropy at long times (r ), between coated and uncoated vesicles (a difference which is more pronounced for cis than for trans parinaric acid), indicates that the presence of the clathrin coat introduces disorder in the surrounding lipids, thus suggesting a possible role of the clathrin in the formation of the pits on the plasma membrane.Abbreviations CVs coated vesicles - UVs uncoated vesicles - TLC thin layer chromatography - DMSO dimethylsulfoxide - DPPC dipalmitoylphosphatidylcholine - cis Pna cis parinaric acid - (9,11,13,15-cis-trans-trans-cis) octadecatetraenoic acid - Trans Pna Trans parinaric acid - (9,11,13,15-all-trans) octadecatetraenoic acid  相似文献   

18.
ATP synthase uses a unique rotational mechanism to convert chemical energy into mechanical energy and back into chemical energy. The helix-turn-helix motif, termed “DELSEED-loop,” in the C-terminal domain of the β subunit was suggested to be involved in coupling between catalysis and rotation. Here, the role of the DELSEED-loop was investigated by functional analysis of mutants of Bacillus PS3 ATP synthase that had 3–7 amino acids within the loop deleted. All mutants were able to catalyze ATP hydrolysis, some at rates several times higher than the wild-type enzyme. In most cases ATP hydrolysis in membrane vesicles generated a transmembrane proton gradient, indicating that hydrolysis occurred via the normal rotational mechanism. Except for two mutants that showed low activity and low abundance in the membrane preparations, the deletion mutants were able to catalyze ATP synthesis. In general, the mutants seemed less well coupled than the wild-type enzyme, to a varying degree. Arrhenius analysis demonstrated that in the mutants fewer bonds had to be rearranged during the rate-limiting catalytic step; the extent of this effect was dependent on the size of the deletion. The results support the idea of a significant involvement of the DELSEED-loop in mechanochemical coupling in ATP synthase. In addition, for two deletion mutants it was possible to prepare an α3β3γ subcomplex and measure nucleotide binding to the catalytic sites. Interestingly, both mutants showed a severely reduced affinity for MgATP at the high affinity site.F1F0-ATP synthase catalyzes the final step of oxidative phosphorylation and photophosphorylation, the synthesis of ATP from ADP and inorganic phosphate. F1F0-ATP synthase consists of the membrane-embedded F0 subcomplex, with, in most bacteria, a subunit composition of ab2c10, and the peripheral F1 subcomplex, with a subunit composition of α3β3γδε. The energy necessary for ATP synthesis is derived from an electrochemical transmembrane proton (or, in some organisms, a sodium ion) gradient. Proton flow down the gradient through F0 is coupled to ATP synthesis on F1 by a unique rotary mechanism. The protons flow through (half) channels at the interface of the a and c subunits, which drives rotation of the ring of c subunits. The c10 ring, together with F1 subunits γ and ε, forms the rotor. Rotation of γ leads to conformational changes in the catalytic nucleotide binding sites on the β subunits, where ADP and Pi are bound. The conformational changes result in the formation and release of ATP. Thus, ATP synthase converts electrochemical energy, the proton gradient, into mechanical energy in the form of subunit rotation and back into chemical energy as ATP. In bacteria, under certain physiological conditions, the process runs in reverse. ATP is hydrolyzed to generate a transmembrane proton gradient, which the bacterium requires for such functions as nutrient import and locomotion (for reviews, see Refs. 16).F1 (or F1-ATPase) has three catalytic nucleotide binding sites located on the β subunits at the interface to the adjacent α subunit. The catalytic sites have pronounced differences in their nucleotide binding affinity. During rotational catalysis, the sites switch their affinities in a synchronized manner; the position of γ determines which catalytic site is the high affinity site (Kd1 in the nanomolar range), which site is the medium affinity site (Kd2 ≈ 1 μm), and which site is the low affinity site (Kd3 ≈ 30–100 μm; see Refs. 7 and 8). In the original crystal structure of bovine mitochondrial F1 (9), one of the three catalytic sites, was filled with the ATP analog AMP-PNP,2 a second was filled with ADP (plus azide) (see Ref. 10), and the third site was empty. Hence, the β subunits are referred to as βTP, βDP, and βE. The occupied β subunits, βTP and βDP, were in a closed conformation, and the empty βE subunit was in an open conformation. The main difference between these two conformations is found in the C-terminal domain. Here, the “DELSEED-loop,” a helix-turn-helix structure containing the conserved DELSEED motif, is in an “up” position when the catalytic site on the respective β subunit is filled with nucleotide and in a “down” position when the site is empty (Fig. 1A). When all three catalytic sites are occupied by nucleotide, the previously open βE subunit assumes an intermediate, half-closed (βHC) conformation. It cannot close completely because of steric clashes with γ (11).Open in a separate windowFIGURE 1.The βDELSEED-loop. A, interaction of the βTP and βE subunits with theγ subunit.β subunits are shown in yellow andγ in blue. The DELSEED-loop (shown in orange, with the DELSEED motif itself in green)of βTP interacts with the C-terminal helixγ and the short helix that runs nearly perpendicular to the rotation axis. The DELSEED-loop of βE makes contact with the convex portion of γ, formed mainly by the N-terminal helix. A nucleotide molecule (shown in stick representation) occupies the catalytic site of βTP, and the subunit is in the closed conformation. The catalytic site on βE is empty, and the subunit is in the open conformation. This figure is based on Protein Data Bank file 1e79 (32). B, deletions in the βDELSEED-loop. The loop was “mutated” in silico to represent the PS3 ATP synthase. The 3–4-residue segments that are removed in the deletion mutants are color-coded as follows: 380LQDI383, pink; 384IAIL387, green; 388GMDE391, yellow; 392LSD394, cyan; 395EDKL398, orange; 399VVHR402, blue. Residues that are the most involved in contacts with γ are labeled. All figures were generated using the program PyMOL (DeLano Scientific, San Carlos, CA).The DELSEED-loop of each of the three β subunits makes contact with the γ subunit. In some cases, these contacts consist of hydrogen bonds or salt bridges between the negatively charged residues of the DELSEED motif and positively charged residues on γ. The interactions of the DELSEED-loop with γ, its movement during catalysis, the conservation of the DELSEED motif (see 1214). Thus, the finding that an AALSAAA mutant in the α3β3γ complex of ATP synthase from the thermophilic Bacillus PS3, where several hydrogen bonds/salt bridges to γ are removed simultaneously, could drive rotation of γ with the same torque as the wild-type enzyme (14) came as a surprise. On the other hand, it seems possible that it is the bulk of the DELSEED-loop, more so than individual interactions, that drives rotation of γ. According to a model favored by several authors (6, 15, 16) (see also Refs. 1719), binding of ATP (or, more precisely, MgATP) to the low affinity catalytic site on βE and the subsequent closure of this site, accompanied by its conversion into the high affinity site, are responsible for driving the large (80–90°) rotation substep during ATP hydrolysis, with the DELSEED-loop acting as a “pushrod.” A recent molecular dynamics (20) study supports this model and implicates mainly the region around several hydrophobic residues upstream of the DELSEED motif (specifically βI386 and βL387)3 as being responsible for making contact with γ during the large rotation substep.

TABLE 1

Conservation of residues in the DELSEED-loop Amino acids found in selected species in the turn region of the DELSEED-loop. Listed are all positions subjected to deletions in the present study. Residue numbers refer to the PS3 enzyme. Consensus annotation: p, polar residue; s, small residue; h, hydrophobic residue; –, negatively charged residue; +, positively charged residue.Open in a separate windowIn the present study, we investigated the function of the DELSEED-loop using an approach less focused on individual residues, by deleting stretches of 3–7 amino acids between positions β380 and β402 of ATP synthase from the thermophilic Bacillus PS3. We analyzed the functional properties of the deletion mutants after expression in Escherichia coli. The mutants showed ATPase activities, which were in some cases surprisingly high, severalfold higher than the activity of the wild-type control. On the other hand, in all cases where ATP synthesis could be measured, the rates where below or equal to those of the wild-type enzyme. In Arrhenius plots, the hydrolysis rates of the mutants were less temperature-dependent than those of wild-type ATP synthase. In those cases where nucleotide binding to the catalytic sites could be tested, the deletion mutants had a much reduced affinity for MgATP at high affinity site 1. The functional role of the DELSEED-loop will be discussed in light of the new information.  相似文献   

19.
The standard assay for iodoperoxidase activity is based on the spectrophotometric detection of triiodide formed during the enzymatic reaction. However, chemical instability of has limited the method to high iodide concentrations and acidic conditions. Here we describe a simple spectrophotometric assay for the determination of iodoperoxidase activities of vanadium haloperoxidases based on the halogenation of thymol blue. The relation between color and chemical entities produced by the vHPO/H2O2/I catalytic system was characterized. The method was extended to bromine and, for the first time, allowed measurement of both iodo- and bromoperoxidase activities using the same assay. The kinetic parameters (Km and kcat) of bromide and iodide for vanadium bromoperoxidase from Ascophyllum nodosum were determined at pH 8.0 from steady-state kinetic analyses. The results are concordant with an ordered two-substrate mechanism. It is proposed that halide selectivity is guided by the chemical reactivity of peroxovanadium intermediate rather than substrate binding. This method is superior to the standard assay, and we believe that it will find applications for the characterization of other vanadium as well as heme haloperoxidases.  相似文献   

20.
Structural changes in theEscherichia coli ATP synthase (ECF1F0) occur as part of catalysis, cooperativity and energy coupling within the complex. The and subunits, two major components of the stalk that links the F1 and F0 parts, are intimately involved in conformational coupling that links catalytic site events in the F1 part with proton pumping through the membrane embedded F0 sector. Movements of the subunit have been observed by electron microscopy, and by cross-linking and fluorescence studies in which reagents are bound to Cys residues introduced at selected sites by mutagenesis. Conformational changes and shifts of the subunit related to changes in nucleotide occupancy of catalytic sites have been followed by similar approaches.  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