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 31 毫秒
1.
Escherichia coli RecBCD is a highly processive DNA helicase involved in double-strand break repair and recombination that possesses two helicase/translocase subunits with opposite translocation directionality (RecB (3′ to 5′) and RecD (5′ to 3′)). RecBCD has been shown to melt out ∼ 5-6 bp upon binding to a blunt-ended duplex DNA in a Mg2+-dependent, but ATP-independent reaction. Here, we examine the binding of E. coli RecBC helicase (minus RecD), also a processive helicase, to duplex DNA ends in the presence and in the absence of Mg2+ in order to determine if RecBC can also melt a duplex DNA end in the absence of ATP. Equilibrium binding of RecBC to DNA substrates with ends possessing pre-formed 3′ and/or 5′ single-stranded (ss)-(dT)n flanking regions (tails) (n ranging from zero to 20 nt) was examined by competition with a fluorescently labeled reference DNA and by isothermal titration calorimetry. The presence of Mg2+ enhances the affinity of RecBC for DNA ends possessing 3′ or 5′-(dT)n ssDNA tails with n < 6 nt, with the relative enhancement decreasing as n increases from zero to six nt. No effect of Mg2+ was observed for either the binding constant or the enthalpy of binding (ΔHobs) for RecBC binding to DNA with ssDNA tail lengths, n ≥ 6 nucleotides. Upon RecBC binding to a blunt duplex DNA end in the presence of Mg2+, at least 4 bp at the duplex end become accessible to KMnO4 attack, consistent with melting of the duplex end. Since Mg2+ has no effect on the affinity or binding enthalpy of RecBC for a DNA end that is fully pre-melted, this suggests that the role of Mg2+ is to overcome a kinetic barrier to melting of the DNA by RecBC and presumably also by RecBCD. These data also provide an accurate estimate (ΔHobs = 8 ± 1 kcal/mol) for the average enthalpy change associated with the melting of a DNA base-pair by RecBC.  相似文献   

2.
An early step in the morphogenesis of the double-stranded DNA (dsDNA) bacteriophage HK97 is the assembly of a precursor shell (prohead I) from 420 copies of a 384-residue subunit (gp5). Although formation of prohead I requires direct participation of gp5 residues 2-103 (Δ-domain), this domain is eliminated by viral protease prior to subsequent shell maturation and DNA packaging. The prohead I Δ-domain is thought to resemble a phage scaffolding protein, by virtue of its highly α-helical secondary structure and a tertiary fold that projects inward from the interior surface of the shell. Here, we employ factor analysis of temperature-dependent Raman spectra to characterize the thermostability of the Δ-domain secondary structure and to quantify the thermodynamic parameters of Δ-domain unfolding. The results are compared for the Δ-domain within the prohead I architecture (in situ) and for a recombinantly expressed 111-residue peptide (in vitro). We find that the α-helicity (∼ 70%), median melting temperature (Tm = 58 °C), enthalpy (ΔHm = 50 ± 5 kcal mol− 1), entropy (ΔSm = 150 ± 10 cal mol− 1 K− 1), and average cooperative melting unit (〈nc〉 ∼ 3.5) of the in situ Δ-domain are altered in vitro, indicating specific interdomain interactions within prohead I. Thus, the in vitro Δ-domain, despite an enhanced helical secondary structure (∼ 90% α-helix), exhibits diminished thermostability (Tm = 40 °C; ΔHm = 27 ± 2 kcal mol− 1; ΔSm = 86 ± 6 cal mol− 1 K− 1) and noncooperative unfolding (〈nc〉 ∼ 1) vis-à-vis the in situ Δ-domain. Temperature-dependent Raman markers of subunit side chains, particularly those of Phe and Trp residues, also confirm different local interactions for the in situ and in vitro Δ-domains. The present results clarify the key role of the gp5 Δ-domain in prohead I architecture by providing direct evidence of domain structure stabilization and interdomain interactions within the assembled shell.  相似文献   

3.
The stability of bovine lactoperoxidase to denaturation by guanidinium–HCl, urea, or high temperature was examined by differential scanning calorimetry (DSC) and tryptophan fluorescence. The calorimetric scans were observed to be dependent on the heating scan rate, indicating that lactoperoxidase stability at temperatures near Tm is controlled by kinetics. The values for the thermal transition, Tm, at slow heating scan rate were 66.8, 61.1, and 47.2 °C in the presence of 0.5, 1, and 2 M guanidinium–HCl, respectively. The extrapolated value for Tm in the absence of guanidinium–HCl is 73.7 °C, compared with 70.2 °C obtained by experiment; a lower experimental value without a denaturant is consistent with distortion of the thermal profile due to aggregation or other irreversible phenomenon. Values for the heat capacity, Cp, at Tm and Ea for the thermal transition decrease under conditions where Tm is lowered. At a given concentration, urea is less effective than guanidinium–HCl in reducing Tm, but urea reduces Cp relatively more. Both fluorescence and DSC indicate that thermally denatured protein is not random coil. A change in fluorescence around 35 °C, which was previously reported for EPR and CD measurements (Boscolo et al. Biochim. Biophys. Acta 1774 (2007) 1164–1172), is not seen by calorimetry, suggesting that a local and not a global change in protein conformation produces this fluorescence change.  相似文献   

4.
5.
6.
Oscar Goñi 《Phytochemistry》2011,72(9):844-854
A 1,3-β-glucanase with potent cryoprotective activity was purified to homogeneity from the mesocarp of CO2-treated cherimoya fruit (Annona cherimola Mill.) stored at low temperature using anion exchange and chromatofocusing chromatography. This protein was characterized as a glycosylated endo-1,3-β-glucanase with a Mr of 22.07 kDa and a pI of 5.25. The hydrolase was active and stable in a broad acidic pH range and it exhibited maximum activity at pH 5.0. It had a low optimum temperature of 35 °C and it retained 40% maximum activity at 5 °C. The purified 1,3-β-glucanase was relatively heat unstable and its activity declined progressively at temperatures above 50 °C. Kinetic studies revealed low kcat (3.10 ± 0.04 s−1) and Km (0.32 ± 0.03 mg ml−1) values, reflecting the intermediate efficiency of the protein in hydrolyzing laminarin. Moreover, a thermodynamic characterization revealed that the purified enzyme displayed a high kcat at both 37 and 5 °C, and a low Ea (6.99 kJ mol−1) within this range of temperatures. In vitro functional studies indicated that the purified 1,3-β-glucanase had no inhibitory effects on Botrytis cinerea hyphal growth and no antifreeze activity, as determined by thermal hysteresis analysis using differential scanning calorimetry. However, a strong cryoprotective activity was observed against freeze-thaw inactivation of lactate dehydrogenase. Indeed, the PD50 was 8.7 μg ml−1 (394 nM), 9.2-fold higher (3.1 on a molar basis) than that of the cryoprotective protein BSA. Together with the observed accumulation of glycine-betaine in CO2-treated cherimoya tissues, these results suggest that 1,3-β-glucanase could be functionally implicated in low temperature-defense mechanism activated by CO2.  相似文献   

7.
F1-ATPase, a water-soluble portion of the enzyme ATP synthase, is a rotary molecular motor driven by ATP hydrolysis. To learn how the kinetics of rotation are regulated, we have investigated the rotational characteristics of a thermophilic F1-ATPase over the temperature range 4-50°C by attaching a polystyrene bead (or bead duplex) to the rotor subunit and observing its rotation under a microscope. The apparent rate of ATP binding estimated at low ATP concentrations increased from 1.2 × 106 M−1 s−1 at 4°C to 4.3 × 107 M−1 s−1 at 40°C, whereas the torque estimated at 2 mM ATP remained around 40 pN·nm over 4-50°C. The rotation was stepwise at 4°C, even at the saturating ATP concentration of 2 mM, indicating the presence of a hitherto unresolved rate-limiting reaction that occurs at ATP-waiting angles. We also measured the ATP hydrolysis activity in bulk solution at 4-65°C. F1-ATPase tends to be inactivated by binding ADP tightly. Both the inactivation and reactivation rates were found to rise sharply with temperature, and above 30°C, equilibrium between the active and inactive forms was reached within 2 s, the majority being inactive. Rapid inactivation at high temperatures is consistent with the physiological role of this enzyme, ATP synthesis, in the thermophile.  相似文献   

8.
Karki B  Maurer D  Jung S 《Bioresource technology》2011,102(11):6522-6528
The effectiveness of several pretreatments [high-power ultrasound, sulfuric acid (H2SO4), sodium hydroxide (NaOH), and ammonium hydroxide (NH3OH)] to enhance glucose production from insoluble fractions recovered from enzyme-assisted aqueous extraction processing of extruded full-fat soybean flakes (FFSF) was investigated. Sonication of the insoluble fraction at 144 μmpp (peak-to-peak) for 30 and 60 s did not improve the saccharification yield. The solid fractions recovered after pretreatment with H2SO4 [1% (w/w), 90 °C, 1.5 h], NaOH [15% (w/w), 65 °C, 17 h], and NH3OH [15% (w/w), 65 °C, 17 h] showed significant lignin degradation, i.e., 81.9%, 71.2%, and 75.4%, respectively, when compared to the control (7.4%). NH3OH pretreatment resulted in the highest saccharification yield (63%) after 48 h of enzymatic saccharification. A treatment combining the extraction and saccharification steps and applied directly to the extruded FFSF, where oil extraction yield and saccharification yield reached 98% and 43%, respectively, was identified.  相似文献   

9.
The thermostability of cellobiohydrolase I Cel7A from Trichoderma reesei was investigated using dynamic light scattering. While the whole enzyme displayed a melting point of 59 °C, the catalytic domain obtained via papain-catalyzed proteolysis was shown to denature at 51 °C and the cellulose-binding domain (with linker attached) melted at 65-66 °C. This variation in individual melting temperatures is proposed to account for the full retention of binding capacity of Cel7A at 50 °C, along with a loss of catalytic activity observed for the catalytic domain alone. Thus, the cellulose-binding domain of Cel7A acts as a thermostabilizing domain for the enzyme. The effect of reducing agents on the protein melting behavior was also investigated.  相似文献   

10.
The binding of a Co(III) complex to the decanucleotide d(CCGAATGAGG)2 containing two pairs of G:A mismatches was studied by 2D-NMR, UV absorption, and molecular modeling. NMR investigations indicate that racemic [Co(phen)2(HPIP)]Cl3 [HPIP = 2-(2-hydroxyphenyl) imidazo [4,5-f][1,10] phenanthroline] binds the decanucleotide by intercalation: the HPIP ligand selectively inserts between the stacked bases from the minor groove at the terminal regions and from the major groove at the sheared region. Further, molecular modeling revealed that the recognition shows strong enantioselectivity: the Λ-isomer preferentially intercalates into the T6G7:A5A4 region from the DNA major groove, while Δ-isomer favors the terminal C1C2:G10G9 region and intercalates from the minor groove. Detailed energy analysis suggests that the steric interaction, especially the electrostatic effect, is the primary determinants of the recognition event. Melting experiments indicate that binding stabilizes the DNA duplex and increases the melting temperature by 9.5 °C. The intrinsic binding constant of the complex to the mismatched duplex was determined to be 3.5 × 105 M−1.  相似文献   

11.
The current study presents a new and novel analysis of heat release signatures measured by a differential scanning calorimeter (DSC) associated with water transport (WT), intracellular ice formation (IIF) and extracellular ice formation (EIF). Correlative cryomicroscopy experiments were also performed to validate the DSC data. The DSC and cryomicroscopy experiments were performed on human dermal fibroblast cells (HDFs) at various cytocrit values (0–0.8) at various cooling rates (0.5–250 °C/min). A comparison of the cryomicroscopy experiments with the DSC analysis show reasonable agreement in the water transport (cellular dehydration) and IIF characteristics between both the techniques with the caveat that IIF measured by DSC lagged that measured by cryomicroscopy. This was ascribed to differences in the techniques (i.e. cell vs. bulk measurement) and the possibility that not all IIF is associated with visual darkening. High and low rates of 0.5 °C/min and 250 °C/min were chosen as HDFs did not exhibit significant IIF or WT at each of these extremes respectively. Analysis of post-thaw viability data suggested that 10 °C/min was the presumptive optimal cooling rate for HDFs and was independent of the cytocrit value. The ratio of measured heat values associated with IIF (qIIF) to the total heat released from both IIF and water transport or from the total cell water content in the sample (qCW) was also found to increase as the cooling rate was increased from 10 to 250 °C/min and was independent of the sample cytocrit value. Taken together, these observations suggest that the proposed analysis is capable of deconvolving water transport and IIF data from the measured DSC latent heat thermograms in cell suspensions during freezing.  相似文献   

12.
Previous investigations have revealed that the boron cluster compound Na2B12H11SH (BSH) is very potent in causing major structural rearrangements of and leakage from phosphatidylcholine liposomes. This somewhat unexpected finding is interesting from a fundamental point of view and may also constitute the basis of future important pharmaceutical/medical applications of BSH. In order to further explore the BSH-lipid interaction, we have studied the effects caused by BSH on dimyristoyl phosphatidylcholine (DMPC) liposomes.Cryo-transmission electron microscopy showed that BSH induces aggregation, membrane rupture and increasing wall thickness of the liposomes. Differential scanning calorimetry revealed a BSH dependent shift of the gel to liquid crystalline phase transition temperature of DMPC. The zeta potential of the liposomes decreases with increasing BSH concentrations, and an apparent dissociation constant of 0.23 mM was found.BSH caused leakage of liposome-encapsulated carboxyfluorescein; leakage was higher at 23 °C (near the phase transition temperature) than at 15 °C and 37 °C. It induced lipid mixing only at very high concentrations.  相似文献   

13.
Sharma PK  Kumar R  Kumar R  Mohammad O  Singh R  Kaur J 《Gene》2012,491(2):264-271
A highly thermostable mutant lipase was generated and characterized. Mutant enzyme demonstrated 144 fold enhanced thermostability over the wild type enzyme at 60 °C. Interestingly, the overall catalytic efficiency (kcat/Km) of mutant was also enhanced (~ 20 folds). Circular dichroism spectroscopy, studied as function of temperature, demonstrated that the mutant lipase retained its secondary structure up to 70-80 °C, whereas wild type protein structure was completely distorted above 35 °C. Additionally, the intrinsic tryptophan fluorescence (a probe for the tertiary structure) also displayed difference in the conformation of two enzymes during temperature dependent unfolding. Furthermore, mutation N355K resulted in extensive H-bonding (Lys355 HZ1OE2 Glu284) with a distance 2.44 Å. In contrast to this, Wt enzyme has not shown such H-bonding interaction.  相似文献   

14.
Deoxyribonucleic acid (DNA) modified gold electrodes are prepared by the dry adsorptive method and the electrochemical behavior of neomycin and the influence of Pb(II) are studied by cyclic voltammetry, chronocoulometry, differential pulse voltammetry. It is found that in 0.01 M phosphate-buffered saline (PBS) buffer solutions (pH 7.3) at DNA/Au electrode neomycin exhibits an irreversible cathodic peak (Ep = 0.489 V), which is more positive and less sensitive compared with that at bare gold electrodes (Ep = 0.423 V). In the presence of Pb(II) the peak shifts toward positive with its height increasing. Moreover, the peak height is linear to neomycin concentration over the range of 0.15-57 μM. The interaction of Pb(II)-neomycin complex with calf thymus DNA is also studied by calculating the binding constants (K) of the Pb(II)-neomycin complex to DNA and binding site size (s) from voltammetric data (1.0 × 107 M−1 and 4 bp, respectively).  相似文献   

15.
The effect of ions on the thermostability and unfolding of Na,K-ATPase from shark salt gland was studied and compared with that of Na,K-ATPase from pig kidney by using differential scanning calorimetry (DSC) and activity assays. In 1 mM histidine at pH 7, the shark enzyme inactivates rapidly at 20 °C, as does the kidney enzyme at 42 °C (but not at 20 °C). Increasing ionic strength by addition of 20 mM histidine, or of 1 mM NaCl or KCl, protects both enzymes against this rapid inactivation. As detected by DSC, the shark enzyme undergoes thermal unfolding at lower temperature (Tm ≈ 45 °C) than does the kidney enzyme (Tm ≈ 55 °C). Both calorimetric endotherms indicate multi-step unfolding, probably associated with different cooperative domains. Whereas the overall heat of unfolding is similar for the kidney enzyme in either 1 mM or 20 mM histidine, components with high mid-point temperatures are lost from the unfolding transition of the shark enzyme in 1 mM histidine, relative to that in 20 mM histidine. This is attributed to partial unfolding of the enzyme due to a high hydrostatic pressure during centrifugation of DSC samples at low ionic strength, which correlates with inactivation measurements. Addition of 10 mM NaCl to shark enzyme in 1 mM histidine protects against inactivation during centrifugation of the DSC sample, but incubation for 1 h at 20 °C prior to addition of NaCl results in loss of components with lower mid-point temperatures within the unfolding transition. Cations at millimolar concentration therefore afford at least two distinct modes of stabilization, likely affecting separate cooperative domains. The different thermal stabilities and denaturation temperatures of the two Na,K-ATPases correlate with the respective physiological temperatures, and may be attributed to the different lipid environments.  相似文献   

16.
In the present study, we report the effects of cooling ejaculated and epididymal rhesus monkey (Macacamulatta) sperm with and without the presence of a cryoprotective agent, glycerol. Water transport data during freezing of ejaculated and epididymal sperm cell suspensions were obtained at a cooling rate of 20 °C/min in the absence of any cryoprotective agents and in the presence of 0.7 M of glycerol, as well. Using previously published values, the macaque sperm cell was modeled as a cylinder of length 73.83 μm with a radius of 0.40 μm and an osmotically inactive cell volume, Vb, of 0.772Vo, where Vo is the isotonic cell volume. This translated to a surface area, SA to initial water volume, WV ratio of ∼22 μm−1. By fitting a model of water transport to the experimentally determined volumetric shrinkage data, the best-fit membrane permeability parameters (reference membrane permeability to water at 0 °C, Lpg or Lpg[cpa] and the activation energy, ELp or ELp[cpa]) were found to range from: Lpg or Lpg[cpa] = 0.0020-0.0029 μm/min-atm; ELp or ELp[cpa]) = 10.6-18.3 kcal/mole. By incorporating these membrane permeability parameters in a recently developed equation (optimal cooling rate, ; where the units of Bopt are °C/min, ELp or ELp[cpa] are kcal/mole, Lpg or Lpg[cpa] are μm/min-atm and SA/WV are μm−1), we determined the optimal rates of freezing macaque sperm to be ∼23 °C/min (ejaculated sperm in the absence of CPAs), ∼29 °C/min (ejaculated sperm in the presence of glycerol), ∼24 °C/min (epididymal sperm in the absence of CPAs) and ∼24 °C/min (epididymal sperm in the presence of glycerol). In conclusion, the subzero water transport response and consequently the subzero water transport parameters are not significantly different between the ejaculated and epididymal macaque spermatozoa under corresponding cooling conditions.  相似文献   

17.
In a previous study we found that 30-40% dimethylsulfoxide induces the active conformation of rabbit muscle pyruvate kinase. Because dimethylsulfoxide is known to perturb structure and function of many proteins, we have explored the effect of trehalose on the kinetics of thermal inactivation and stability of pyruvate kinase; this is because trehalose, in contrast to dimethyl sulfoxide, is totally excluded from the hydration shell of proteins. The results show that 600 mM trehalose inhibits the activity of pyruvate kinase by about 20% at 25 °C, however, trehalose protects pyruvate kinase from thermal inactivation at 60 °C, increases the Tmapp of unfolding by 7.2 °C, induces a more compact state, and stabilizes its tetrameric structure. The inactivation process is irreversible due to the formation of protein aggregates. Trehalose diminishes the rate of formation of intermediates with propensity to aggregate, but does not affect the extent of aggregation. Remarkably, trehalose affects the aggregation process by inducing aggregates with amyloid-like characteristics.  相似文献   

18.
Control of choline oxidation in rat kidney mitochondria   总被引:1,自引:0,他引:1  
Niaobh O'Donoghue 《BBA》2009,1787(9):1135-1087
Choline is a quaternary amino cationic organic alcohol that is oxidized to betaine in liver and kidney mitochondria. Betaine acts as an intracellular organic osmolyte in the medulla of the kidney. Evidence is provided that kidney mitochondria have a choline transporter in their inner membrane. The transporter has a Km of 173 ± 64 μM and a Vmax of 0.4 ± 0.1 nmol/min/mg mitochondrial protein (at 10 °C). Uptake of choline is not coupled to betaine efflux. Transporter activity demonstrates a dependence on membrane potential and choline transport is inhibited by hemicholinium-3. Steady-state oxygen consumption due to choline oxidation in kidney mitochondria was measurable at 37 °C (125 ± 6 pmolO2/min/mg mitochondrial protein), in the absence of other mitochondrial electron transport chain substrates and the choline transporter was shown to be the major site of control (96 ± 4%) over choline oxidation flux in isolated kidney mitochondria. We conclude that the choline transporter in rat kidney mitochondria is the major site of control over the production of the organic osmolyte, betaine.  相似文献   

19.
We have adapted the thermal shift assay to measure the ligand binding properties of the herpes simplex virus-1 single-strand DNA binding protein, ICP8. By measuring SYPRO Orange fluorescence in microtiter plates using a fluorescence-enabled thermal cycler, we have quantified the effects of oligonucleotide ligands on the melting temperature of ICP8. We found that single-stranded oligomers raise the melting temperature of ICP8 in a length- and concentration-dependent manner, ranging from 1 °C for (dT)5 to a maximum of 9 °C with oligomers ?10 nucleotides, with an apparent Kd of <1 μM for (dT)20. Specifically, the results indicate that ICP8 is capable of interacting with oligomers as short as 5 nucleotides. Moreover, the observed increases in melting temperature of up to 9 °C, indicates that single-strand DNA binding significantly stabilizes the structure of ICP8. This assay may be applied to investigate the ligand binding proteins of other single-strand DNA binding proteins and used as a high-throughput screen to identify compounds with therapeutic potential that inhibit single-strand DNA binding. As proof of concept, the single-strand DNA binding agent ciprofloxacin reduces the ligand induced stabilization of the melting temperature of ICP8 in a dose-dependent manner.  相似文献   

20.
Phytate, the major source of phosphorus in seeds, exists as a complex with different metal ions. Alkaline phytases are known to dephosphorylate phytate complexed with calcium ions in contrast to acid phytases that act only on phytic acid. A recombinant alkaline phytase from Bacillus sp. MD2 has been purified and characterized with respect to the effect of divalent metal ions on the enzyme activity and stability. The presence of Ca2+ on both the enzyme and the substrate is required for optimal activity and stability. Replacing Ca2+ with Ba2+, Mn2+, Mg2+ and Sr2+ in the phytase resulted in the expression of > 90% of the maximal activity with calcium-phytate as the substrate, while Fe2+ and Zn2+ rendered the enzyme inactive. On the other hand, the calcium loaded phytase showed significant activity (60%) with sodium phytate and lower activity (17-20%) with phytate complexed with only Mg2+, Sn2+ and Sr2+, respectively. On replacing Ca2+ on both the enzyme and the substrate with other metal ions, about 20% of the maximal phytase activity was obtained only with Mg2+ and Sr2+, respectively. Only Ca2+ resulted in a marked increase in the melting temperature (Tm) of the enzyme by 12-21 °C, while Ba2+, Mn2+, Sr2+ or Cu2+ resulted in a modest (2-3.5 °C) increase in Tm. In the presence of 1-5 mM Ca2+, the optimum temperature of the phytase activity was increased from 40 °C to 70 °C, while optimum pH of the enzyme shifted by 0.4-1 pH unit towards the acidic region.  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