首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
The optical rotation phenomena exhibited by a citrate-extracted fraction of ichthyocol (from carp swim bladder), as well as by the parent gelatin derived therefrom, have been studied. Dispersion data for all cases follow a single-term Drude equation, but the variations with state are adequately expressed by simple reference to changes in [alpha](D) as follows:- 1. The native collagen fraction, dispersed in 0.15 M citrate buffer at pH 3.7 in the cold (11 degrees C.), yields a high negative specific rotation, [alpha](D), near -350 degrees . 2. During equilibration at 40 degrees C., which causes conversion to a monodisperse parent gelatin, the rotation drops to about -110 degrees . 3. Gelation at 2 degrees C. results in a partial regain of rotation to around -290 degrees . This mutarotation is reversible, depending on temperature. 4. In the range 0.02 to 0.28 per cent the native ichthyocol and the warm gelatin solutions show little concentration dependence, but with the cold gelatin solutions the specific rotation increases with concentration. Gelatin films formed by cold evaporation yield high specific rotation (ca. -620 degrees ), but those formed by hot evaporation retain low optical activity. 5. Since this same collagen-gelatin system has been investigated physicochemically, it is possible to relate molecular changes to the observed variations in optical rotation. Conclusions are similar to those of Robinson (1953), who studied other gelatins: high negative rotation is believed related to a native collagen polypeptide configuration, herein specified as helical (from x-ray diffraction considerations) and destroyed by heating. The possible roles of intermolecular interactions and of prevalent pyrrolidine constituents in influencing the helical configuration and optical activity are discussed.  相似文献   

2.
1. Purified citrate-extracted ichthyocol obtained from carp swim bladders has been further characterized with respect to its content of certain amino acids and carbohydrate substances. 2. The degree of solubilization or dispersion of ichthyocol by solutions of certain salts maintained in the range of neutral pH and at a temperature of 0-2 degrees C. has been determined. 3. While a number of salts of monovalent cations had no significant solubilizing effects on ichthyocol, ammonium chloride in a concentration of 1 M did cause solution of the protein. 4. Sodium thiosulfate in a range of concentrations caused the solubilization of ichthyocol but was most effective in an intermediate concentration of 0.25 M. 5. Several salts of divalent cations, in particular the chlorides of calcium, magnesium, and barium, and magnesium thiosulfate in concentrations ranging from 0.3 to 1 M caused the immediate and complete solubilization of the ichthyocol. 6. Solutions of ichthyocol in calcium chloride, magnesium chloride, and sodium thiosulfate buffered or adjusted to pH 7.0, were studied with respect to intrinsic viscosity of the protein, optical rotation, ultracentrifugal sedimentation, and reconstitution into fibers. It was found in each case that the original characteristics of the collagen, as determined previously in acid solution, were maintained when the protein was dissolved in salt solutions of neutral pH. No evidence of denaturation or gelatinization could be found when ichthyocol was solubilized under the stated conditions. 7. Collagen in neutral solution with sodium thiosulfate, calcium chloride, or magnesium chloride was not attacked by trypsin as determined viscometrically at 20.0 degrees C., but was rapidly degraded by a purified bacterial collagenase.  相似文献   

3.
A new series of 1β-methylcarbapenems 1a–i bearing isoxazoloethenyl groups on the pyrrolidine ring has been prepared and evaluated for in vitro antibacterial activity and stability to DHP-I. Most compounds showed excellent antibacterial activity and high stability to DHP-I superior to that of meropenem. Of these new carbapenems, 1a,b,h exhibited the best combination of antibacterial activity and DHP-I stability.  相似文献   

4.
F1-ATPase is a rotary molecular motor in which the γ-subunit rotates against the α3β3 cylinder. The unitary γ-rotation is a 120° step comprising 80 and 40° substeps, each of these initiated by ATP binding and ADP release and by ATP hydrolysis and inorganic phosphate release, respectively. In our previous study on γ-rotation at low temperatures, a highly temperature-sensitive (TS) reaction step of F1-ATPase from thermophilic Bacillus PS3 was found below 9 °C as an intervening pause before the 80° substep at the same angle for ATP binding and ADP release. However, it remains unclear as to which reaction step the TS reaction corresponds. In this study, we found that the mutant F1(βE190D) from thermophilic Bacillus PS3 showed a clear pause of the TS reaction below 18 °C. In an attempt to identify the catalytic state of the TS reaction, the rotation of the hybrid F1, carrying a single copy of βE190D, was observed at 18 °C. The hybrid F1 showed a pause of the TS reaction at the same angle as for the ATP binding of the incorporated βE190D, although kinetic analysis revealed that the TS reaction is not the ATP binding step. These findings suggest that the TS reaction is a structural rearrangement of β before or after ATP binding.F1-ATPase (F1)2 is an ATP-driven rotary motor protein. The subunit composition of the bacterial F1-ATPase is α3β3γδϵ, and the minimum complex of F1-ATPase as a rotary motor is α3β3γ subcomplex. This motor protein forms the FoF1-ATP synthase complex by binding to another rotary motor, namely, Fo, which is driven by the proton flux resulting from the proton motive force across the membranes (14). Under physiological conditions, where the proton motive force is sufficiently large, Fo forcibly rotates F1-ATPase in the reverse direction of F1-ATPase, leading the reverse reaction of ATP hydrolysis, i.e. ATP synthesis from ADP and inorganic phosphate (Pi). When the proton motive force diminishes or F1 is isolated from Fo, F1-ATPase hydrolyzes ATP to rotate the γ-subunit against the α3β3 stator ring in the counterclockwise direction as viewed from the Fo side (5). The catalytic sites are located at the interface of the α- and β-subunits, predominantly on the β-subunit (6). Each β-subunit carries out a single turnover of ATP hydrolysis during the γ-rotation of 360° following the common catalytic reaction pathway, whereas they are 120° different in the catalytic phase. In this manner, the three β-subunits undergo different reaction steps of ATP hydrolysis upon each rotational step. The rotary motion of the γ-subunit has been demonstrated by biochemical (7) and spectroscopic methods (8) and directly proved in single-molecule observation studies (5).Since the establishment of the single-molecule rotation assay, the chemomechanical coupling scheme of F1 has been studied extensively by resolving the rotation into discrete steps. The stepping rotation was first observed under an ATP-limiting condition where F1 makes discrete 120° steps upon ATP binding (9). Then, high speed imaging of the rotation with a small probe of low friction was performed, which revealed that the 120° step comprises 80 and 40° substeps, each initiated by ATP binding, and two unknown consecutive reactions, respectively (10). This finding necessitated the identification of the two reactions that trigger the 40° substep. Hence, the rotation assay was performed using a mutant, namely F1(βE190D), and a slowly hydrolyzed ATP analog, namely ATPγS (11). Glutamate 190 of the β-subunit of F1, derived from thermophilic Bacillus PS3 and the corresponding glutamates from other F1-ATPases (Glu-181 of F1 from Escherichia coli and Glu-188 of F1 from bovine mitochondria), has been identified as one of the most critical catalytic residues for ATP hydrolysis (6, 1215). When this glutamate was substituted with aspartic acid, which has a shorter side chain than that of glutamate, the ATP cleavage step of F1 was drastically slowed. In the rotation assay, this mutant showed a distinct long pause before the 40° substep. ATPγS also caused a long pause before the 40° substep. These observations established that the 40° substep is initiated by hydrolysis. Accordingly, the pause angles before the 80 and 40° substeps are referred to as to the binding angle and the catalytic angle, respectively. Then, the rotation assay was performed in the presence of a high amount of Pi in the solution. It was shown that Pi rebinding caused the long pause at the catalytic angle, suggesting that Pi is released before the 40° substep (16).However, the reaction scheme of F1 cannot be established by simply assigning each reaction step to either the binding angle or the catalytic angle, because each reaction step must be assigned to one of the three binding or catalytic angles when considering the 360° cyclic reaction scheme of each β-subunit. Direct information about the timing of ADP release was obtained by simultaneous imaging of fluorescently labeled nucleotides and γ rotation, which showed that each β retains ADP until the γ rotates 240° after binding of the nucleotide as ATP and releases ADP between 240 and 320° (16, 17). Another powerful approach is the use of a hybrid F1 carrying a mutant β that causes a characteristic pause during the rotation. In a previous study, the hybrid F1 carrying a single copy of β(E190D), α3β2β(E190D)γ, showed a distinct pause caused by the slow hydrolysis of β(E190D) at +200° from the ATP binding angle of the mutant β (18). From this observation, it was confirmed that each β executes the chemical cleavage of the bound ATP at +200° from the angle where the ATP binds to β. The asymmetric feature of the pause of the hybrid F1 was also utilized in other experiments as a marker in the rotational trajectory to correlate the rotational angle and the conformational state of β (19) or to determine the state of F1 in the crystal structures as the pausing state at catalytic angle (20).Recently, we have found a new reaction intermediate of F1 rotation as a clear intervening pause before the 80° substep in the rotation assay below 9 °C (21). Furuike et al. (22) also observed the TS reaction in a high speed imaging experiment. The rate constant of this reaction was remarkably sensitive to temperature, giving a Q10 factor around 19. When ADP was added to solution, the pause before the 80° substep was prolonged, whereas the solution Pi caused a longer pause before the 40° substep (21). Although this result can be explained by assuming that the temperature-sensitive (TS) reaction is ADP release, it was not decisive for the identification of the TS reaction.In this study, we found that the mutant F1(βE190D) also exhibits the distinct pause of the TS reaction but at a higher temperature than for the wild-type F1, i.e. at 18 °C. This feature was advantageous in identifying the angle position of the TS reaction in the catalytic cycle for each β-subunit coupled with the 360° rotation. Taking advantage of the feature of the hybrid F1, we analyzed the rotational behavior of the hybrid F1 at 18 °C in order to assign the angle position of the TS reaction in the catalytic cycle of the 360° rotation, and we have shown that the TS reaction is not directly involved in the ADP release but in some conformational rearrangement before or after ATP binding step.  相似文献   

5.
Rat fibrinogen was purified from rat plasma by using lysine–Sepharose chromatography, repeated precipitation with 25%-satd. (NH4)2SO4 and gel chromatography on Sepharose 6B. To minimize proteolytic activity, rats were injected intravenously with Trasylol before bleeding and the collected blood was treated with Trasylol and di-isopropyl phosphorofluoridate. A 95%-clottable preparation was obtained in 70–75% yield; it proved to be free of factor XIII and plasminogen. It showed a single band on sodium dodecyl sulphate/polyacrylamide-gel electrophoresis and on disc electrophoresis in 8m-urea. Alanine was the only detectable N-terminal amino acid. After reduction and modification of the thiol groups, the material could be separated into three distinct chains (Aα, Bβ and γ) by pore-limit polyacrylamide slab-gel electrophoresis in the presence of sodium dodecyl sulphate. The amino acid compositions of the whole fibrinogen and of the separated modified chains were determined. The molecular weights were 61000, 58000 and 51000 for Aα-, Bβ- and γ-chains respectively. Our results for the chains are in contrast with previous reports on rat fibrinogen [Bouma & Fuller (1975) J. Biol. Chem. 250, 4678–4683; Stemberger & Jilek (1976) Thromb. Res. 9, 657–660], in which no separation between Aα- and Bβ-chains was achieved on sodium dodecyl sulphate/polyacrylamide-gel electrophoresis for 3h. Evidence is presented that this is probably due to Aα-chain degradation as a result of incomplete inhibition of proteolytic enzymes during the purification. Complete inhibition of proteolytic activities is essential in all steps of the present purification procedure.  相似文献   

6.
Characteristics of thermal denaturation of pea legumin and a product of its limited proteolysis with trypsin – legumin-T, in a wide range of NaCl concentrations have bean measured by means of differential scanning microcalorimetry. By the increase of NaCl concentration, the number of cooperative units (domains) increases from 1 per one polypeptide chain to 2 for legumin and 1.8 for legumin-T. Deconvolution of denaturation peaks have revealed up to three peaks, which were ascribed to the dissociation of protein macromolecules to subunits and the unfolding of - and β-polypeptide chains. The analysis of experimental data based on some assumptions showed that the splitting of C-termini of -chains, which are not constituents of cooperative domains, in the course of limited trypsinolysis results in destabilization of the quaternary structure of legumin and loosening of -chains, as well as decrease of the temperatures of their maximum stability.  相似文献   

7.
Crespo MD  Rubini M 《PloS one》2011,6(5):e19425

Background

Many strategies have been employed to increase the conformational stability of proteins. The use of 4-substituted proline analogs capable to induce pre-organization in target proteins is an attractive tool to deliver an additional conformational stability without perturbing the overall protein structure. Both, peptides and proteins containing 4-fluorinated proline derivatives can be stabilized by forcing the pyrrolidine ring in its favored puckering conformation. The fluorinated pyrrolidine rings of proline can preferably stabilize either a Cγ-exo or a Cγ-endo ring pucker in dependence of proline chirality (4R/4S) in a complex protein structure. To examine whether this rational strategy can be generally used for protein stabilization, we have chosen human ubiquitin as a model protein which contains three proline residues displaying Cγ-exo puckering.

Methodology/Principal Findings

While (2S,4R)-4-fluoroproline ((4R)-FPro) containing ubiquitinin can be expressed in related auxotrophic Escherichia coli strain, all attempts to incorporate (2S,4S)-4-fluoroproline ((4S)-FPro) failed. Our results indicate that (4R)-FPro is favoring the Cγ-exo conformation present in the wild type structure and stabilizes the protein structure due to a pre-organization effect. This was confirmed by thermal and guanidinium chloride-induced denaturation profile analyses, where we observed an increase in stability of −4.71 kJ·mol−1 in the case of (4R)-FPro containing ubiquitin ((4R)-FPro-ub) compared to wild type ubiquitin (wt-ub). Expectedly, activity assays revealed that (4R)-FPro-ub retained the full biological activity compared to wt-ub.

Conclusions/Significance

The results fully confirm the general applicability of incorporating fluoroproline derivatives for improving protein stability. In general, a rational design strategy that enforces the natural occurring proline puckering conformation can be used to stabilize the desired target protein.  相似文献   

8.
ATP synthases from coupling membranes are complex rotary motors that convert the energy of proton gradients across coupling membranes into the chemical potential of the β-γ anhydride bond of ATP. Proton movement within the ring of c subunits localized in the F0-sector drives γ and ɛ rotation within the F1α3β3 catalytic core where substrates are bound and products are released. An external stalk composed of homodimeric subunits b2 in Escherichia coli or heterodimeric bb′ in photosynthetic synthases connects F0 subunit a with F1 subunits δ and most likely α. The external stalk resists rotation, and is of interest both functionally and structurally. Hypotheses that the external stalk contributes to the overall efficiency of the reaction through elastic coupling of rotational substeps, and that stalks form staggered, right-handed coiled coils, are investigated here. We report on different structures that accommodate heptad discontinuities with either local or global underwinding. Analyses of the knob-and-hole packing of the E. coli b2 and Synechocystis bb′ stalks strongly support the possibility that these proteins can adopt conventional left-handed coiled coils.  相似文献   

9.
10.
F1-ATPase is the water-soluble part of ATP synthase and is an ATP-driven rotary molecular motor that rotates the rotary shaft against the surrounding stator ring, hydrolyzing ATP. Although the mechanochemical coupling mechanism of F1-ATPase has been well studied, the molecular details of individual reaction steps remain unclear. In this study, we conducted a single-molecule rotation assay of F1 from thermophilic bacteria under various pressures from 0.1 to 140 MPa. Even at 140 MPa, F1 actively rotated with regular 120° steps in a counterclockwise direction, showing high conformational stability and retention of native properties. Rotational torque was also not affected. However, high hydrostatic pressure induced a distinct intervening pause at the ATP-binding angles during continuous rotation. The pause was observed under both ATP-limiting and ATP-saturating conditions, suggesting that F1 has two pressure-sensitive reactions, one of which is evidently ATP binding. The rotation assay using a mutant F1(βE190D) suggested that the other pressure-sensitive reaction occurs at the same angle at which ATP binding occurs. The activation volumes were determined from the pressure dependence of the rate constants to be +100 Å3 and +88 Å3 for ATP binding and the other pressure-sensitive reaction, respectively. These results are discussed in relation to recent single-molecule studies of F1 and pressure-induced protein unfolding.  相似文献   

11.
The biosynthesis of the pyrrolidine ring of nicotine has been studied using short-term steady-state exposures of Nicotiana glutinosa seedlings to 14CO2. The pyrrolidine ring of the labeled nicotine has been degraded in a systematic manner to ascertain the radioactivity at each carbon, and a new method has been developed for obtaining C-2′ with complete radiochemical integrity. Some of the labeling patterns obtained were symmetrical while others were clearly unsymmetrical. The duality of the labeling patterns found in these 14CO2 biosyntheses, together with other data on pyrrolidine ring biosynthesis which are critically examined, is best rationalized by postulating two biosynthetic pathways for formation of the pyrrolidine ring, one involving a symmetrical precursor and the other an unsymmetrical one.  相似文献   

12.
The temperature-dependent rotation of F1-ATPase γ subunit was observed in Vmax conditions at low viscous drag using a 60-nm gold bead (Nakanishi-Matsui, M., Kashiwagi, S., Hosokawa, H., Cipriano, D. J., Dunn, S. D., Wada, Y., and Futai, M. (2006) J. Biol. Chem. 281, 4126–4131). The Arrhenius slopes of the speed of the individual 120° steps and reciprocal of the pause length between rotation steps were very similar, indicating a flat energy pathway followed by the rotationally coupled catalytic cycle. In contrast, the Arrhenius slope of the reciprocal pause length of the γM23K mutant F1 was significantly increased, whereas that of the rotation rate was similar to wild type. The effects of the rotor γM23K substitution and the counteracting effects of βE381D mutation in the interacting stator subunits demonstrate that the rotor-stator interactions play critical roles in the utilization of stored elastic energy. The γM23K enzyme must overcome an abrupt activation energy barrier, forcing it onto a less favored pathway that results in uncoupling catalysis from rotation.F-ATPase (FoF1), consisting of the catalytic sector F13β3γδϵ) and the transmembrane proton transport sector Fo (ab2c10), synthesizes or hydrolyzes ATP coupled with proton transport (for reviews, see Ref. 16). As Abrahams et al. (7) discovered in the first high resolution x-ray structure, a critical feature of the F1-ATPase is the inherent asymmetry of the three β subunits in different conformations, βTP, βDP, and βE, referring to the nucleotide bound in each catalytic site, ATP, ADP, or empty, respectively. A rotational mechanism has been firmly established mostly based on direct observation in single molecule experiments of the behavior of the rotor complex ϵγc10, relative to the stator complex α3β3δab2 (reviewed in Ref. 1). ATP hydrolysis-dependent rotation of the γ and ϵ subunits in purified bacterial F1 (8, 9), the ϵγc10 complex in detergent solubilized FoF1 (1013), and the ϵγc10 complexin FoF1 in lipid bilayers (14) were shown experimentally by single molecule observations using fluorescent actin filament as a probe. Relative rotation of the single copy Fo a subunit was also shown in F0F1, which was immobilized through the ring of ∼10 c subunits, suggesting that the rotor and stator are interchangeable mechanical units (14). ATP synthesis by F-ATPase is believed to follow the reverse mechanism of ATP hydrolysis because mechanically induced rotation of the γ subunit in immobilized F1 in the presence of ADP and Pi results in net ATP synthesis (15, 16). There remain many questions about the mechanism of coupling between catalysis and transport via mechanical rotation. In particular, the mechanism of coupling H+ transport to rotation of the subunit c10 ring is still not well understood (4).In contrast, there is considerably more information on the mechanism of coupling catalysis to γ and ϵ subunit rotation. Observations of γ subunit rotation in the catalytic F1 sector are consistent with Boyer''s binding change model (17); thus coupling between the chemistry and rotation can be assessed by studies of the soluble F1, and these findings relate to the mechanism of the entire ATP synthase complex. The γ subunit rotates relative to the α3β3 hexamer in distinct 120° steps. A 120° rotation step consisting of pause and rotation substeps appears to correspond to the hydrolysis of one ATP, assuming that three ATP molecules are hydrolyzed per 360° revolution (18). Additional pauses observed at low ATP concentrations are attributed to the “ATP waiting” dwell (19). Yasuda et al. (19) and Shimabukuro et al. (20) further resolved that each 120° step occurred in two substeps: an 80° substep whose onset was dependent upon the Mg·ATP concentration, and a 40° substep, which was not affected by substrate concentration (19). The pause before the 80° substep, the ATP waiting dwell became shorter with increasing [Mg·ATP]. In contrast, the pause duration before the 40° rotation step was modulated by the slow hydrolysis rate of ATPγS2 or by the catalytic site mutant βE190D (in the Bacillus PS3 F1), which was found to significantly increase the length of the catalytic dwell (20). These data together indicate that the dwell before the 40° step is the “catalytic dwell” (20) and defines the order of the substeps during the 120° rotation step observed in high Mg·ATP concentrations (21).In this paper, we address the question of when the rate-limiting step of steady state catalysis occurs, with respect to the rotational behavior. Pre-steady state analysis of the burst kinetics of ATP hydrolysis at nearly Vmax conditions demonstrated that the rate-limiting transition state occurs after the reversible hydrolysis/synthesis step and before release of phosphate (Pi) (22, 23). The rate-limiting step is likely associated with a rotation step because a γ-β cross-linked enzyme is still able to undergo the initial ATP hydrolysis, but the rotation-impeded enzyme is unable to release Pi (23). Significantly, the kinetics of steady state hydrolysis can only be assessed when the Mg·ATP concentration is high enough to fill all three catalytic sites. The only model consistent with these data is one that involves all three catalytic sites. During each 120° catalytic cycle, one site binds ATP, a different site carries out reversible hydrolysis/synthesis, and the third site releases product Pi and ADP (22, 23).Steady state analyses, which take advantage of a particular γ subunit mutation γM23K (24), are consistent with this model. Replacement of the conserved γMet-23 with lysine causes an uncoupling between catalysis and γ subunit rotation caused by altered interactions between γ and β subunits (25). Importantly, Al-Shawi and Nakamoto (26) and Al-Shawi et al. (25, 27) found that the γM23K mutation strongly affected the rate-limiting transition state of steady state ATP hydrolysis and ATP synthesis. The slope of the Arrhenius plots and thus the energy of activation were significantly increased in the mutant enzyme. Several second site suppressor mutations, mostly in the γ subunit (28, 29) but also in the β subunits (30, 31), were genetically identified because they restored coupled ATP synthesis. Significantly, all were in the γ-β interface. Thermodynamic analyses found that the second site suppressors generally compensated for the primary γM23K mutations by reducing the increased activation energy (25, 27, 31). Although most of the second site mutations were found distant from the γM23K site, the x-ray crystal structures (7) suggested that γM23K may directly interact with conserved βGlu-381. As expected, replacement of βGlu-381 with aspartate also suppressed the uncoupling effects of γM23K (31).To identify the rate-limiting transition state step in the rotational behavior, we analyzed the temperature dependence of the γM23K mutant in Vmax conditions observed in single molecule experiments. Interestingly, direct observation of this mutant using the micron-length actin filaments did not detect differences in the rotation behavior at room temperature (9). In contrast, we find in the data presented here that there is dramatic effect of the mutation on the temperature dependence of the length of the catalytic dwell or pause between the 120° rotation steps. This is likely because of two factors: first, we used a bead small enough not to invoke a drag on the rotation (32), and second, the temperature dependence of the rate of the rotation steps is critical for the analyses of the mechanism.  相似文献   

13.
Glycosidases and glycosyl transferases fall into two major mechanistic classes; those that hydrolyse the glycosidic bond with retention of anomeric configuration and those that do so with inversion. There are, however, two classes of transferases: those that use nucleotide phosphosugars (NP-sugar-dependent) and those that simply transglycosylate between oligosaccharides or polysaccharides (transglycosylases). The latter are mechanistically similar to retaining glycosidases while the mechanisms of NP-sugar-dependent transferases are far from clear.

Retaining glycosidases and the transglycosylases employ a mechanism involving a covalent glycosyl–enzyme intermediate formed and hydrolysed with acid/base catalytic assistance via oxocarbenium ion-like transition states. This intermediate has been trapped on glycosidases in two distinct ways, either by modification of the substrate through fluorination, or of the enzyme through mutation of key residues. A third method has been developed for trapping the intermediate on transglycosylases involving the use of incompetent substrates that allow formation of the intermediate, but prohibit its transfer as a consequence of their acceptor hydroxyl group being removed.

Three-dimensional structures of several of these glycosyl–enzyme complexes, along with those of Michaelis complexes, have been determined through X-ray crystallographic analysis, revealing the identities of important amino acid residues involved in catalysis. In particular they reveal the involvement of the carbonyl oxygen of the catalytic nucleophile in strong hydrogen bonding to the sugar 2-hydroxyl for the β-retainers or in interactions with the ring oxygen for -retainers. The glucose ring in the −1 (cleavage) site in the intermediates formed on several cellulases and a β-glucosidase adopts a normal 4C1 chair conformation. By contrast the xylose ring at this site in a xylanase is substantially distorted into a 2,5B boat conformation, an observation that bears significant stereoelectronic implications. Substantial distortion is also observed in the substrate upon binding to several β-glycosidases, this time to a 1S3 skew boat conformation. Much less distortion is seen in the substrate bound on an -transglycosylase.

Finally an efficient catalyst for synthesis, but not hydrolysis, of glycosidic bonds has been generated by mutation of the glutamic acid catalytic nucleophile of a β-glucosidase to an alanine. When used with -glucosyl fluoride as a glycosyl donor, along with a suitable acceptor, oligosaccharides up to five sugars in length have been made with yields of up to 90% on individual steps. These new enzymes have been named Glycosynthases.  相似文献   


14.
Protein C inhibitor (PCI) is a serpin type of serine protease inhibitor that is found in many tissues and fluids in human, including blood plasma, seminal plasma and urine. This inhibitor displays an unusually broad protease specificity compared with other serpins. Previous studies have shown that the N-glycan(s) and the NH2-terminus affect some blood-related functions of PCI. In this study, we have for the first time determined the N-glycan profile of seminal plasma PCI, by mass spectrometry. The N-glycan structures differed markedly compared with those of both blood-derived and urinary PCI, providing evidence that the N-glycans of PCI are expressed in a tissue-specific manner. The most abundant structure (m/z 2592.9) had a composition of Fuc3Hex5HexNAc4, consistent with a core fucosylated bi-antennary glycan with terminal Lewisx. A major serine protease in semen, prostate specific antigen (PSA), was used to evaluate the effects of N-glycans and the NH2-terminus on a PCI function related to the reproductive tract. Second-order rate constants for PSA inhibition by PCI were 4.3±0.2 and 4.1±0.5 M−1s−1 for the natural full-length PCI and a form lacking six amino acids at the NH2-terminus, respectively, whereas these constants were 4.8±0.1 and 29±7 M−1s−1 for the corresponding PNGase F-treated forms. The 7–8-fold higher rate constants obtained when both the N-glycans and the NH2-terminus had been removed suggest that these structures jointly affect the rate of PSA inhibition, presumably by together hindering conformational changes of PCI required to bind to the catalytic pocket of PSA.  相似文献   

15.
During ATP hydrolysis by F1-ATPase subunit γ rotates in a hydrophobic bearing, formed by the N-terminal ends of the stator subunits (αβ)3. If the penultimate residue at the α-helical C-terminal end of subunit γ is artificially cross-linked (via an engineered disulfide bridge) with the bearing, the rotary function of F1 persists. This observation has been tentatively interpreted by the unfolding of the α-helix and swiveling rotation in some dihedral angles between lower residues. Here, we screened the domain between rotor and bearing where an artificial disulfide bridge did not impair the rotary ATPase activity. We newly engineered three mutants with double cysteines farther away from the C-terminus of subunit γ, while the results of three further mutants were published before. We found ATPase and rotary activity for mutants with cross-links in the single α-helical, C-terminal portion of subunit γ (from γ285 to γ276 in E. coli), and virtually no activity when the cross-link was placed farther down, where the C-terminal α-helix meets its N-terminal counterpart to form a supposedly stable coiled coil. In conclusion, only the C-terminal singular α-helix is prone to unwinding and can form a swivel joint, whereas the coiled coil portion seems to resist the enzyme''s torque.  相似文献   

16.
ATP synthase uses a unique rotary mechanism to couple ATP synthesis and hydrolysis to transmembrane proton translocation. As part of the synthesis mechanism, the torque of the rotor has to be converted into conformational rearrangements of the catalytic binding sites on the stator to allow synthesis and release of ATP. The γ subunit of the rotor, which plays a central role in the energy conversion, consists of two long helices inside the central cavity of the stator cylinder plus a globular portion outside the cylinder. Here, we show that the N-terminal helix alone is able to fulfill the function of full-length γ in ATP synthesis as long as it connects to the rest of the rotor. This connection can occur via the ϵ subunit. No direct contact between γ and the c ring seems to be required. In addition, the results indicate that the ϵ subunit of the rotor exists in two different conformations during ATP synthesis and ATP hydrolysis.F1Fo-ATP synthase is responsible for the bulk of ATP synthesis from ADP and Pi in most organisms. F1Fo-ATP synthase consists of the membrane-embedded Fo subcomplex with, in most bacteria, a subunit composition of ab2cn (with n = 10–15) 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, sodium ion) gradient. Proton flow, down the gradient, through Fo is coupled to ATP synthesis on F1 by a unique rotary mechanism. The protons flow through channels at the interface of a and c subunits, which drives rotation of the ring of c subunits. The cn 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 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 can run in reverse. ATP is hydrolyzed to generate a transmembrane proton gradient that 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 with 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, 8). Only the high affinity site is catalytically active (9). In the original crystal structure of bovine mitochondrial F1 (10), one of the three catalytic sites was filled with the ATP analog AMPPNP,3 a second one with ADP (plus azide; see Ref. 11), and the third site was empty. Hence, the β subunits are referred to as βTP, βDP, and βE, respectively. The high affinity site is located on the βTP subunit (12).The coupling process between ATP synthesis or hydrolysis on β and rotation of γ is not yet fully understood on a residue level. A number of point mutations at the interfaces between α or β and γ and between γ, ϵ, and c have been described that result in varying degrees of uncoupling (1317). Some mutations at these interfaces were found that abolish ATP synthesis or hydrolysis activity or both (1820). Considering the pronounced effect of these point mutations, some of which were even conservative substitutions, it came as a surprise when it was recently shown that an “axle-less” γ, consisting just of the globular portion, with the portions of the N- and C-terminal helices that reach into the α3β3 cylinder removed, displayed ATP-driven rotation in the correct direction (21).Some reports have implicated the ϵ subunit (corresponding to the δ subunit in the mitochondrial enzyme) as being involved in coupling (15, 2225). It was shown that ϵ exists in different conformations that vary in the folding and positioning of the C-terminal domain. The x-ray structure of the mitochondrial enzyme (26) shows the two helices of the C-terminal domain folded back on each other like a hairpin and positioned close to the interface between γ and the c ring (“down” conformation). In the crystal structure of a γϵ complex from Escherichia coli the hairpin is unfolded (27); when integrated into F1 or F1Fo, the C terminus would reach “up,” coming close to the DELSEED-loop of the α and/or β subunits. While in this up conformation the angle between both helices of the C-terminal domain of ϵ is ∼90°, it has been postulated that this domain might also exist in a fully extended up conformation, stretching close to the N terminus of γ, with helical regions replacing the turn between the two helices (28). Fixing ϵ in either up conformation by cross-linking to γ has been shown to impair ATP hydrolysis but not synthesis. Freezing ϵ in the down position inhibited neither reaction (29, 30).Here, we report a finding that is arguably just as surprising as the rotation of an axle-less γ. In ATP synthase from the thermophilic bacterium Bacillus PS3, enzymes with γ subunit constructs where the globular domain and the C-terminal helix were eliminated, consisting of just the N-terminal 35 or 42 residues, TF1Fo(γQ36stop)4 and TF1Fo(γP43stop), were able to catalyze significant rates of ATP synthesis. According to the crystal structure (26), the shorter of the two γ constructs should not make any contact either with c or with ϵ in the down conformation. Thus, the fact that ATP synthesis was observed suggests that ϵ in an up conformation forms a complex with the truncated γ, which is able to catalyze ATP synthesis. Indeed, when the γQ36stop truncation was combined with an ϵ truncation where the C-terminal domain was removed, ATP synthesis was abolished. The functions of the γ and ϵ subunits will be discussed in light of the new findings.  相似文献   

17.
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   

18.
The sequences of the C-terminal octadecapeptides obtained by cyanogen bromide cleavage of the γ-chains of myeloma proteins of the four subclasses, and a urinary heavy-chain-disease protein, have been determined. Although the sequences were markedly homologous, unique replacements were identified that distinguished between the γ2b, γ2c and γ2d subclasses. The data are in accord with the postulated existence of four genetic loci or cistrons, these having arisen by the process of gene duplication.  相似文献   

19.
β-Carrageenan, essentially devoid of ester sulfate, was isolated from the hot aqueous extracts of alkali-modified Eucheuma gelatinae, Eucheuma speciosa, and Endocladia muricatum by precipitating the more anionic moieties with a quaternary ammonium salt, isolating the fractions that did not precipitate, then treating these with an anion-exchange cellulose. The β-carrageenan was characterized by chemical analysis, optical rotation, and NMR. Gelling was found to be ion-independent, with Tg = 31–33°C and Tm = 63–70°C. Specific optical rotations of the isolated β-carrageenan samples were more positive than the κ-, λ-, and ι-carrageenans with which they were compared, while agarose, its stereoisomer, exhibited a negative specific rotation. Electrophoresis gels made from β-carrageenan were used to separate DNA fragments which exhibited faster migration than on an agarose gel of comparable concentration, indicating that β-carrageenan has a less restrictive pore structure.  相似文献   

20.
Gatti DL 《PloS one》2012,7(1):e30079

Background

The first line of defense by bacteria against β-lactam antibiotics is the expression of β-lactamases, which cleave the amide bond of the β-lactam ring. In the reaction of biapenem inactivation by B2 metallo β-lactamases (MβLs), after the β-lactam ring is opened, the carboxyl group generated by the hydrolytic process and the hydroxyethyl group (common to all carbapenems) rotate around the C5–C6 bond, assuming a new position that allows a proton transfer from the hydroxyethyl group to C2, and a nucleophilic attack on C3 by the oxygen atom of the same side-chain. This process leads to the formation of a bicyclic compound, as originally observed in the X-ray structure of the metallo β-lactamase CphA in complex with product.

Methodology/Principal Findings

QM/MM and metadynamics simulations of the post-hydrolysis steps in solution and in the enzyme reveal that while the rotation of the hydroxyethyl group can occur in solution or in the enzyme active site, formation of the bicyclic compound occurs primarily in solution, after which the final product binds back to the enzyme. The calculations also suggest that the rotation and cyclization steps can occur at a rate comparable to that observed experimentally for the enzymatic inactivation of biapenem only if the hydrolysis reaction leaves the N4 nitrogen of the β-lactam ring unprotonated.

Conclusions/Significance

The calculations support the existence of a common mechanism (in which ionized N4 is the leaving group) for carbapenems hydrolysis in all MβLs, and suggest a possible revision of mechanisms for B2 MβLs in which the cleavage of the β-lactam ring is associated with or immediately followed by protonation of N4. The study also indicates that the bicyclic derivative of biapenem has significant affinity for B2 MβLs, and that it may be possible to obtain clinically effective inhibitors of these enzymes by modification of this lead compound.  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