首页 | 本学科首页   官方微博 | 高级检索  
文章检索
  按 检索   检索词:      
出版年份:   被引次数:   他引次数: 提示:输入*表示无穷大
  收费全文   34篇
  免费   0篇
  2015年   1篇
  2014年   5篇
  2013年   1篇
  2012年   6篇
  2011年   2篇
  2010年   1篇
  2009年   5篇
  2008年   3篇
  2007年   2篇
  2006年   1篇
  2005年   3篇
  2004年   1篇
  2003年   3篇
排序方式: 共有34条查询结果,搜索用时 31 毫秒
1.
There is a mounting body of evidence to suggest that enzyme motions are linked to function, although the design of informative experiments aiming to evaluate how this motion facilitates reaction chemistry is challenging. For the family of diflavin reductase enzymes, typified by cytochrome P450 reductase, accumulating evidence suggests that electron transfer is somehow coupled to large-scale conformational change and that protein motions gate the electron transfer chemistry. These ideas have emerged from a variety of experimental approaches, including structural biology methods (i.e. X-ray crystallography, electron paramagnetic/NMR spectroscopies and solution X-ray scattering) and advanced spectroscopic techniques that have employed the use of variable pressure kinetic methodologies, together with solvent perturbation studies (i.e. ionic strength, deuteration and viscosity). Here, we offer a personal perspective on the importance of motions to electron transfer in the cytochrome P450 reductase family of enzymes, drawing on the detailed insight that can be obtained by combining these multiple structural and biophysical approaches.  相似文献   
2.
Air-borne microbial contamination of surfaces in a UK dental clinic   总被引:1,自引:0,他引:1  
Little is known about the number, type, or antibiotic resistance profiles, of air-borne microbes present in hospital settings yet such information is important in designing effective measures to reduce cross-infection. In this study settle plates were used to identify and quantify the air-borne microbes present in a dental clinic. All isolates were identified to species level using partial 16S ribosomal RNA gene sequencing and their susceptibility to ampicillin, chloramphenicol, erythromycin, gentamicin, penicillin, tetracycline or vancomycin was performed. The mean numbers of viable bacteria detected for each sampling occasion during periods of clinical activity and in the absence of such activity were 21.9x10(2 )cfu/m(2)/h and 2.3x10(2 )cfu/m(2)/h respectively. One hundred ninety-three distinct colony morphotypes, comprising 73 species, were isolated during the study and 48% of these were resistant to at least one antibiotic. The mean numbers of different morphotypes detected per sampling occasion were 14.3 and 5 during periods of clinical activity and inactivity respectively. Propionibacterium acnes, Micrococcus luteus and Staphylococcus epidermidis were frequently isolated regardless of whether any clinical activities were taking place. These findings highlight the importance of preventing surfaces from becoming reservoirs of antibiotic-resistant bacteria and thereby contributing to cross-infection in the dental clinic.  相似文献   
3.
Phytochromes are dimeric photoreceptors that regulate a range of responses in plants and microorganisms through interconversion of red light-absorbing (Pr) and far-red light-absorbing (Pfr) states. Photoconversion between these states is initiated by light-driven isomerization of a bilin cofactor, which triggers protein structural change. The extent of this change, and how light-driven structural changes in the N-terminal photosensory region are transmitted to the C-terminal regulatory domain to initiate the signalling cascade, is unknown. We have used pulsed electron-electron double resonance (PELDOR) spectroscopy to identify multiple structural transitions in a phytochrome from Synechocystis sp. PCC6803 (Cph1) by measuring distances between nitroxide labels introduced into the protein. We show that monomers in the Cph1 dimer are aligned in a parallel ‘head-to-head’ arrangement and that photoconversion between the Pr and Pfr forms involves conformational change in both the N- and C-terminal domains of the protein. Cryo-trapping and kinetic measurements were used to probe the extent and temporal properties of protein motions for individual steps during photoconversion of Cph1. Formation of the primary photoproduct Lumi-R is not affected by changes in solvent viscosity and dielectric constant. Lumi-R formation occurs at cryogenic temperatures, consistent with their being no major structural reorganization of Cph1 during primary photoproduct formation. All remaining steps in the formation of the Pfr state are affected by solvent viscosity and dielectric constant and occur only at elevated temperatures, implying involvement of a series of long-range solvent-coupled conformational changes in Cph1. We show that signalling is achieved through ultrafast photoisomerization where localized structural change in the GAF domain is transmitted and amplified to cause larger-scale and slower conformational change in the PHY and histidine kinase domains. This hierarchy of timescales and extent of structural change orientates the histidine kinase domain to elicit the desired light-activated biological response.  相似文献   
4.
The light-activated enzyme NADPH-protochlorophyllide oxidoreductase (POR) catalyzes the trans addition of hydrogen across the C-17–C-18 double bond of protochlorophyllide (Pchlide), a key step in chlorophyll biosynthesis. Similar to other members of the short chain alcohol dehydrogenase/reductase family of enzymes, POR contains a conserved Tyr and Lys residue in the enzyme active site, which are implicated in a proposed reaction mechanism involving proton transfer from the Tyr hydoxyl group to Pchlide. We have analyzed a number of POR variant enzymes altered in these conserved residues using a combination of steady-state turnover, laser photoexcitation studies, and low temperature fluorescence spectroscopy. None of the mutations completely abolished catalytic activity. We demonstrate their importance to catalysis by defining multiple roles in the overall reaction pathway. Mutation of either residue impairs formation of the ground state ternary enzyme-substrate complex, pointing to a key role in substrate binding. By analyzing the most active variant (Y193F), we show that Tyr-193 participates in proton transfer to Pchlide and stabilizes the Pchlide excited state, enabling hydride transfer from NADPH to Pchilde. Thus, in addition to confirming the probable identity of the proton donor in Pchlide reduction, our work defines additional roles for these residues in facilitating hydride transfer through stabilization of the ground and excited states of the ternary enzyme complex.The light-driven enzyme protochlorophyllide oxidoreductase (POR)3 (EC 1.3.1.33) catalyzes the trans addition of hydrogen across the C-17–C-18 double bond of the chlorophyll precursor protochlorophyllide (Pchlide) (see Fig. 1A) (1). This reaction is a key step in the synthesis of chlorophyll and leads to profound changes in the morphological development of photosynthetic organisms through modification and reorganization of plastid membranes (2, 3). In addition to POR, nonflowering land plants, algae, and cyanobacteria possess a light-independent Pchlide reductase, which consists of three separate subunits and allows these organisms to produce chlorophyllide in the dark (4). Together with DNA photolyase (5), POR is one of only two enzymes studied so far that exhibit a direct, natural requirement for light and because mixing strategies are no longer required to initiate the reaction, it is possible to trigger catalysis on very fast time scales and at cryogenic temperatures. Consequently, POR has proven to be an excellent model system for studying the role of protein dynamics in driving enzyme catalysis (1).Open in a separate windowFIGURE 1.The light-driven reduction of Pchlide. A, the trans addition of hydrogen across the C-17–C-18 double bond of Pchlide to form chlorophyllide (Chlide) in the chlorophyll biosynthesis pathway is catalyzed by the light-driven enzyme POR. B, shown is a three-dimensional model of the POR-catalyzed reaction based on the structural homology model of POR (26) and the proposed mechanism of hydride and proton transfer (8). Upon activation by light, a hydride is transferred to the C-17 position of Pchlide from the pro-S face of NADPH (shown in yellow), and the proton at the C-18 position is derived from Tyr-193 (shown in cyan). The conserved Lys-197 residue (shown in magenta) is proposed to decrease the pKa of the Tyr to facilitate the proton transfer reaction.In the POR catalytic cycle, a ternary enzyme-NADPH-Pchlide complex is formed. Following light activation of this complex, a hydride ion is transferred from the pro-S face of NADPH to the C-17 atom of Pchlide (6, 7). The valence of the C-18 atom is satisfied by proton transfer, which is suggested to originate from an active site tyrosine residue (8). The catalytic cycle of POR has been analyzed through the trapping of intermediates at cryogenic temperatures. Following the initial light-driven reaction (9), there are a series of subsequent (slower) dark reactions (10, 11). The light-driven step involves hydride transfer from NADPH to form a charge transfer complex, which then facilitates protonation of the pigment intermediate during the first of the “dark” reactions (12). Moreover, through laser activation of catalysis, we have shown that both of these H-transfer reactions proceed by quantum mechanical tunneling coupled to motions in the enzyme-substrate complex on the submicrosecond time scale (13). The final dark steps in the reaction cycle involve a series of ordered product release and cofactor binding steps linked to conformational changes in the enzyme (10, 11, 14). Ultrafast measurements have uncovered spectral changes on the picosecond timescale that are likely to represent conformational changes prior to Pchlide reduction (1519). Previous excitation of POR with a laser pulse leads to a more efficient conformation of the active site and an enhancement in the catalytic efficiency of the enzyme (18).POR is a member of a large family of enzymes known as short chain dehydrogenases/reductases (SDR). These are single domain NAD(P)+- or NAD(P)H-binding oxidoreductases that exist generally as dimers or tetramers (20). A number of SDR enzymes (e.g. carbonyl reductase, alcohol dehydrogenase, and dihydrofolate reductase) have been good model systems for studying the dynamics linked to enzyme catalysis (2123). This family of enzymes has been amenable to studies of biological H-tunneling (2426), and in particular the unique light-activated properties of POR make it an excellent system for studying mechanisms of H-transfer and dynamics in this family of enzymes.Structures of several SDR family members are available, and these have enabled the construction of a homology model of POR from Synechocystis (27). This model comprises a central parallel β-sheet of seven β-strands, surrounded by nine α-helices, with an additional unique 33-residue insertion between the fifth and sixth β-sheets. The NADPH cofactor binds within the N-terminal region of the enzyme, which contains a common nucleotide-binding motif with a tight βαβ fold, termed the Rossmann fold (27). Importantly, a Tyr and a Lys residue are both absolutely conserved throughout all members of the SDR family and are critical for catalysis in a number of enzymes (2831). A common mechanism has been proposed for this group of enzymes, involving a Tyr-X-X-X-Lys motif. The Lys residue in this motif is presumed to facilitate proton donation from the Tyr hydroxyl group to substrate through favorable perturbation of the hydroxyl group pKa (8, 31). In POR, multiple turnover assays have also indicated that these Tyr and Lys residues are important for activity (8, 32, 33), leading to a proposed mechanism that involves proton transfer from the conserved Tyr residue to the C-18 position of Pchlide (8) (Fig. 1B). The close proximity of the Lys residue is thought to allow the deprotonation step to occur at physiological pH by lowering the apparent pKa of the phenolic group of the Tyr (8). However, confirmation of the exact role of these conserved residues has been compromised by the limited levels of activity observed in previous studies of the variant enzymes (8, 32, 33), and a detailed evaluation of the role of the active site Tyr and Lys residues on the chemical steps (i.e. hydride and proton transfer) has not been reported. We address this deficiency in the current work by analyzing a number of site-specific mutant forms altered at Tyr-193 and Lys-197 in a thermophilic POR from Thermosynechococcus elongatus BP-1. This was achieved using steady-state (multiple turnover) and laser photoexcitation (single turnover) methods and by trapping transient reaction intermediates by fluorescence spectroscopy performed at cryogenic temperatures.  相似文献   
5.
The reduction of nitrite (NO2) into nitric oxide (NO), catalyzed by nitrite reductase, is an important reaction in the denitrification pathway. In this study, the catalytic mechanism of the copper-containing nitrite reductase from Alcaligenes xylosoxidans (AxNiR) has been studied using single and multiple turnover experiments at pH 7.0 and is shown to involve two protons. A novel steady-state assay was developed, in which deoxyhemoglobin was employed as an NO scavenger. A moderate solvent kinetic isotope effect (SKIE) of 1.3 ± 0.1 indicated the involvement of one protonation to the rate-limiting catalytic step. Laser photoexcitation experiments have been used to obtain single turnover data in H2O and D2O, which report on steps kinetically linked to inter-copper electron transfer (ET). In the absence of nitrite, a normal SKIE of ∼1.33 ± 0.05 was obtained, suggesting a protonation event that is kinetically linked to ET in substrate-free AxNiR. A nitrite titration gave a normal hyperbolic behavior for the deuterated sample. However, in H2O an unusual decrease in rate was observed at low nitrite concentrations followed by a subsequent acceleration in rate at nitrite concentrations of >10 mm. As a consequence, the observed ET process was faster in D2O than in H2O above 0.1 mm nitrite, resulting in an inverted SKIE, which featured a significant dependence on the substrate concentration with a minimum value of ∼0.61 ± 0.02 between 3 and 10 mm. Our work provides the first experimental demonstration of proton-coupled electron transfer in both the resting and substrate-bound AxNiR, and two protons were found to be involved in turnover.Denitrification is an anaerobic respiration pathway found in bacteria, archaea, and fungi, in which ATP synthesis is coupled to the sequential reduction of nitrate (NO3) and nitrite (NO2) (NO3 → NO2 → NO → N2O → N2) (13).3 The first committed step in this reaction cascade is the formation of gaseous NO by nitrite reductase (NiR), the key enzyme of this pathway. Two distinct classes of periplasmic NiR are found in denitrifying bacteria, one containing cd1 hemes as prosthetic groups (46) and the other utilizing two copper centers to catalyze the one-electron reduction of nitrite (7). Copper-containing NiRs are divided into two main groups according to the color of their oxidized type 1 copper center (T1Cu), with shades ranging from blue to green (3, 7). NiR from Alcaligenes xylosoxidans subsp. xylosoxidans (NCIMB 11015, AxNiR), which is analyzed in this study, is a member of the blue CuNiR group. The blue and green subclasses show a high degree of sequence similarity (70%) (8) and have similar trimeric structures with each monomer (∼36.5 kDa in AxNiR) consisting of two greek key β-barrel cupredoxin-like motifs as well as one long and two short α-helical regions (7, 9).Each NiR monomer contains two copper-binding sites per catalytic unit. One is a T1Cu center, which receives electrons from a physiological redox partner protein and is buried 7 Å beneath the protein surface (10), and the other copper is a type 2 center (T2Cu), constituting the catalytically active substrate-binding site (11). The physiological electron donor for the blue NiRs are the small copper protein azurin (14 kDa) (7) and cytochrome c551 (7, 12, 13). The T1Cu, which is responsible for the color of NiR, serves as the electron delivery center and is coordinated by two histidine residues as well as one cysteine and one methionine residue. The catalytic T2Cu, which like all T2Cu centers has very weak optical bands, is ligated to three His residues and an H2O/OH ligand in the resting state. This H2O/OH ligand is held in place by hydrogen bonds to the active site residues, Asp-92 (AxNiR numbering) and His-249, and gets displaced by the substrate during catalytic turnover (14). The T2Cu is located at the base of a 13–14-Å substrate access channel at the interface of two monomers with one of the three His residues being part of the adjacent subunit (15, 16). The two copper centers are connected by a 12.6-Å covalent bridge provided by the T1Cu-coordinating Cys and by one of the T2Cu His ligands (17, 18). This linkage has been suggested to constitute the electron transfer (ET) pathway from the T1Cu center to the catalytically active T2Cu center via 11 covalent bonds (19).Intramolecular ET from T1- to T2Cu has been extensively examined using pulse radiolysis studies (7, 1924). In a variety of NiR species, ET could be measured, both in the presence and absence of substrate, with observed ET rate constants (kET(obs)) ranging from ∼150 to ∼2000 s−1. According to the Marcus semi-classical ET theory (25), the redox potentials (E0, redox midpoint potential at pH 7.0) of the copper centers affect both the thermodynamic equilibrium and the ET kinetics. In the absence of substrate, the difference in the redox potentials has been found to be insignificant at pH 7 (E0 (T1Cu) ∼240 mV and E0 (T2Cu) ∼230 mV (20)), implying a thermodynamically equal electron distribution between the two metal centers. From an enzymatic point of view, however, approaching this equilibrium position on such a fast time scale (≥150 s−1) is unfavorable in the absence of substrate, as NiR has been shown to form an inactive species with a reduced T2Cu that is devoid of the H2O/OH ligand and unable to bind nitrite (26, 27). Substrate binding has been proposed to induce a favorable shift in the T2Cu redox potential, which would be expected to result in an accelerated ET compared with the substrate-free reaction (7, 16, 25, 2730). However, kET(obs) values in AxgNiR (GIFU1051) have been demonstrated to be lower in the nitrite-bound than in the substrate-free enzyme between pH 7.7 and 5.5 (21). Below pH 5.5, the ET rate constants were observed to be similar in the nitrite-free and -bound enzyme (21).In addition to changes in the redox potentials and thus in the driving force of the ET reaction, several structural changes in the redox centers have been reported as a result of substrate binding, which may also influence the inter-copper ET rate by changing the reorganization energy (16, 25, 30, 31). These rearrangements include subtle changes in the Cys-His bridge linking T1- and T2Cu (32) and conformational transitions of the catalytically relevant active site residue Asp-92 (see below and Ref. 29). Moreover, the presence of nitrite has been postulated to be relayed to the T1Cu site via the so-called substrate sensor loop (via His-94, Asp-92, and His-89 in AxNiR), thereby triggering ET to the T2Cu (19, 27, 29, 32). The tight coupling of ET to the presence of substrate has been argued to prevent the formation of a deactivated enzyme species with a prematurely reduced T2Cu (14, 16, 19, 26, 27, 33). In accordance with such a feedback mechanism, in a combined crystallographic and single-crystal spectroscopic study, inter-copper ET could only be detected in crystals where nitrite was bound to the T2Cu site, whereas in the absence of substrate no such ET was observed (34). This finding, however, contradicts the pulse radiolysis results at room temperature (see above), and the apparent discrepancy between solution studies and x-ray crystallographic data collected at cryogenic temperature remains to be resolved.The one-electron reduction of nitrite to NO involves two protons according to the chemical net equation NO2 + 2H+ + e → NO + H2O, if the T2Cu is ligated by an H2O molecule in the resting state rather than an OH ion. Although the exact enzymatic mechanism is still somewhat controversial (35, 36), one suggested reaction sequence is given in Scheme 1. The potential participation of active site residues in catalyzing the proton transfer (PT) steps has been investigated by studying the pH dependence of NiR under steady-state conditions as well as by pulse radiolysis. The trends obtained for kcat and kET(obs), are similar with pH optima between 5.2 and 6, indicating the involvement of two amino acid residues (21, 22, 37). Asp-92 and His-249 have been proposed as acid-base catalysts (18, 21, 22, 28, 38), and the abrupt drop in rates at increasing pH may indicate that OH can act as a competitive inhibitor for nitrite (39). The relevance of these active site residues, however, as well as the timing of the two protonation steps is still a matter of debate (35, 40, 41).4Open in a separate windowSCHEME 1.A potential reaction mechanism proposed for CuNiRs. Adapted from Ref. 36. Nitrite is shown to bind to the oxidized T2Cu as nitrous acid, thus involving the first protonation step. It coordinates to the oxidized T2Cu center in a bidentate fashion. Following inter-copper ET yielding a reduced T2Cu center, the initially deprotonated Asp-92 accepts a proton, which is subsequently transferred to the substrate. His-249 may be a potential source of this second proton. PT and ET reactions may be reversible and they may be concerted rather than sequential as suggested by the arrows. See text for further information.There are no experimental studies that have been aimed at directly examining the kinetic coupling of PT and ET steps in AxNiR. In this study of the blue AxNiR, our aims were to gain further insight into the mechanism of nitrite reduction by combining multiple turnover experiments with laser photoexcitation studies to measure the (single turnover) inter-copper ET. An extensive analysis of the solvent kinetic isotope effect (SKIE) has been employed as a means of determining whether solvent-exchangeable protons and/or water molecules play a rate-limiting role in the catalytic turnover and/or in inter-copper ET.  相似文献   
6.
Protein dynamics are crucial for realizing the catalytic power of enzymes, but how enzymes have evolved to achieve catalysis is unknown. The light-activated enzyme protochlorophyllide oxidoreductase (POR) catalyzes sequential hydride and proton transfers in the photoexcited and ground states, respectively, and is an excellent system for relating the effects of motions to catalysis. Here, we have used the temperature dependence of isotope effects and solvent viscosity measurements to analyze the dynamics coupled to the hydride and proton transfer steps in three cyanobacterial PORs and a related plant enzyme. We have related the dynamic profiles of each enzyme to their evolutionary origin. Motions coupled to light-driven hydride transfer are conserved across all POR enzymes, but those linked to thermally activated proton transfer are variable. Cyanobacterial PORs require complex and solvent-coupled dynamic networks to optimize the proton donor-acceptor distance, but evolutionary pressures appear to have minimized such networks in plant PORs. POR from Gloeobacter violaceus has features of both the cyanobacterial and plant enzymes, suggesting that the dynamic properties have been optimized during the evolution of POR. We infer that the differing trajectories in optimizing a catalytic structure are related to the stringency of the chemistry catalyzed and define a functional adaptation in which active site chemistry is protected from the dynamic effects of distal mutations that might otherwise impact negatively on enzyme catalysis.  相似文献   
7.
In chlorophyll biosynthesis protochlorophyllide reductase (POR) catalyzes the light-driven reduction of protochlorophyllide (Pchlide) to chlorophyllide, providing a rare opportunity to trap and characterize catalytic intermediates at low temperatures. Moreover, the presence of a chlorophyll-like molecule allows the use of EPR, electron nuclear double resonance, and Stark spectroscopies, previously used for the analysis of photosynthetic systems, to follow catalytic events in the active site of POR. Different models involving the formation of either radical species or charge transfer complexes have been proposed for the initial photochemical step, which forms a nonfluorescent intermediate absorbing at 696 nm (A696). Our EPR data show that the concentration of the radical species formed in the initial photochemical step is not stoichiometric with conversion of substrate. Instead, a large Stark effect, indicative of charge transfer character, is associated with A696. Two components were required to fit the Stark data, providing clear evidence that charge transfer complexes are formed during the initial photochemistry. The temperature dependences of both A696 formation and NADPH oxidation are identical, and we propose that formation of the A696 state involves hydride transfer from NADPH to form a charge transfer complex. A catalytic mechanism of POR is suggested in which Pchlide absorbs a photon, creating a transient charge separation across the C-17-C-18 double bond, which promotes ultrafast hydride transfer from the pro-S face of NADPH to the C-17 of Pchlide. The resulting A696 charge transfer intermediate facilitates transfer of a proton to the C-18 of Pchlide during the subsequent first "dark" reaction.  相似文献   
8.
Dental plaque from 76 children without amalgam restorations was screened for bacteria resistant to mercuric chloride. Seventy-one per cent of the children harboured mercury-resistant oral bacteria and the median percentage of the total oral microflora resistant to mercuric chloride was 0.007% (range 0-5.3%). Eighty-seven mercury-resistant bacteria were isolated and 86% of these were streptococci with Streptococcus mitis predominating. Sixty per cent of the mercury-resistant isolates were also resistant to at least one of the four antibiotics tested (penicillin, ampicillin, erythromycin and tetracycline) with resistance to tetracycline (40% of isolates) most frequently encountered.  相似文献   
9.
We demonstrate that photoexcitation of NAD(P)H reduces heme iron of Mycobacterium tuberculosis P450s CYP121 and CYP51B1 on the microsecond time scale. Rates of formation for the ferrous-carbonmonoxy (Fe(II)-CO) complex were determined across a range of coenzyme/CO concentrations. CYP121 reaction transients were biphasic. A hyperbolic dependence on CO concentration was observed, consistent with the presence of a CO binding site in ferric CYP121. CYP51B1 absorption transients for Fe(II)-CO complex formation were monophasic. The reaction rate was second order with respect to [CO], suggesting the absence of a CO-binding site in ferric CYP51B1. In the absence of CO, heme iron reduction by photoexcited NAD(P)H is fast ( approximately 10,000-11,000 s(-1)) with both P450s. For CYP121, transients revealed initial production of the thiolate-coordinated (P450) complex (absorbance maximum at 448 nm), followed by a slower phase reporting partial conversion to the thiol-coordinated P420 species (at 420 nm). The slow phase amplitude increased at lower pH values, consistent with heme cysteinate protonation underlying the transition. Thus, CO binding occurs to the thiolate-coordinated ferrous form prior to cysteinate protonation. For CYP51B1, slow conversions of both the ferrous/Fe(II)-CO forms to species with spectral maxima at 423/421.5 nm occurred following photoexcitation in the absence/presence of CO. This reflected conversion from ferrous thiolate- to thiol-coordinated forms in both cases, indicating instability of the thiolate-coordinated ferrous CYP51B1. CYP121 Fe(II)-CO complex pH titrations revealed reversible spectral transitions between P450 and P420 forms. Our data provide strong evidence for P420 formation linked to reversible heme thiolate protonation, and demonstrate key differences in heme chemistry and CO binding for CYP121 and CYP51B1.  相似文献   
10.
The latter stages of the catalytic cycle of the light-driven enzyme, protochlorophyllide oxidoreductase, have been investigated using novel laser photoexcitation methods. The formation of the ternary product complex was initiated with a 6-ns laser pulse, which allowed the product release steps to be kinetically accessed for the first time. Subsequent absorbance changes associated with the release of the NADP+ and chlorophyllide products from the enzyme could be followed on a millisecond timescale. This has facilitated a detailed kinetic and thermodynamic characterization for the interconversion of all the various bound and unbound product species. Initially, NADP+ is released from the enzyme in a biphasic process with rate constants of 1210 and 237 s(-1). The rates of both phases show a significant dependence on the viscosity of the solvent and become considerably slower at higher glycerol concentrations. The fast phase of this process exhibits no dependence on NADP+ concentration, suggesting that conformational changes are required prior to NADP+ release. Following NADP+ release, the NADPH rebinds to the enzyme with a maximum rate constant of approximately 72 s(-1). At elevated temperatures (>298 K) chlorophyllide is released from the enzyme to yield the free product with a maximum rate constant of 20 s(-1). The temperature dependencies of the rates of each of these steps were measured, and enthalpies and entropies of activation were calculated using the Eyring equation. A comprehensive kinetic and thermodynamic scheme for these final stages of the reaction mechanism is presented.  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