首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
A double mutant of CuA azurin was prepared in which both bridging cysteine thiolate ligands of the binuclear CuA center were replaced by serine. The copper binding properties of this protein were investigated, and shown to be pH dependent. At lower pH (5.2 ± 0.1), the protein binds one copper per protein molecule as demonstrated by electrospray ionization mass spectrometry. Copper titrations resulted in electronic absorptions at 730 nm (peak) and ca. 330 nm (shoulder) in the UV-Vis spectrum. EPR data show a four line pattern with hyperfine A = 150 G and g and g values 2.32 and 2.03, characteristic of a type II (T2) copper. Superhyperfines to two nitrogen atoms were also observed. At higher pH (8.5 ± 0.1), the protein binds upto two copper atoms per protein molecule, and copper titrations exhibit a blue transition at 595 nm in the UV-Vis spectrum. The EPR data are consistent with two monomeric sites very similar to one another having hyperfines A = 182 and 150 G, g = 2.24 and 2.22 and a similar g value of 2.01. These results indicate that both bridging cysteines play a critical role in the CuA center, and replacing them with serines is not enough to maintain the symmetrical diamond core structure or the characteristic electronic and functional properties of the CuA center.  相似文献   

2.
The occurrence of summer mortalities of the commercially important Pacific oyster, Crassostrea gigas, has increased in recent years. These mortality events occur during the late summer when water temperatures are at their highest. Many theories have been proposed concerning the causes including reproductive stress, environmental stress, disease, or synergistic interactions of these factors. C. gigas are grown intertidally and are exposed to the air (emersed) for hours at a time. These organisms can experience extreme changes in temperature during the course of a day. An oyster closed during emersion depletes the oxygen stores to near zero within the shell and builds up CO2 causing a decrease in tissue pH. The focus of this study is to determine the respiratory (pH, Po2, Pco2 and total CO2) and immune responses of oysters exposed to air at normal seasonal temperatures, and to determine whether these stresses associated with emersion inhibit the immune system of the oyster and contribute to the summer mortalities. The respiratory variables of the hemolymph of oysters submerged at 18 °C (pH = 7.52 ± 0.04 S.E.M., Po2 = 7.09 ± 0.53 S.E.M. kPa and Pco2 = 0.20 ± 0.03 S.E.M. kPa) varied significantly from oysters emersed for four hours at 22°C (pH = 7.11 ± 0.03 S.E.M., Po2 = 3.83 ± 0.15 S.E.M. kPa, Pco2 = 0.36 ± 0.03 S.E.M. kPa) and those emersed for four hours at 30 °C (pH = 6.84 ± 0.02 S.E.M., Po2 = 3.10 ± 0.12 S.E.M. kPa, Pco2 = 1.31 ± 0.06 S.E.M. kPa). The ability of hemocytes to kill the bacterium Vibrio campbellii was assessed using an in vitro assay to generate a killing index. There was no significant difference in the killing index between pH treatment groups (p = 0.856): at pH 7.6 killing index = 50.2% ± 2.33 S.E.M., at pH 6.6 killing index = 52.3% ± 3.67 S.E.M.. Temperature was the only factor to significantly affect the killing indices among temperature and oxygen treatment groups. The killing index was lowest (29.3% ± 3.25 S.E.M.) at 30 °C and 7% oxygen, simulating in vivo oxygen pressure in well-aerated conditions and 30 °C and 3% oxygen, simulating in vivo oxygen pressure in hypoxia (30.5% ± 3.25 S.E.M.), compared with the index in 7% oxygen at low temperature (18 °C) (44.4% ± 4.50 S.E.M.) or compared with low oxygen (3%) at low temperature (18 °C) (39.7% ± 2.51 S.E.M.). The seasonal and diurnal rise in temperature may, therefore, be an important factor contributing to summer mortalities of C. gigas.  相似文献   

3.
A dispersion of melting temperatures at pH 5.3 for individual residues of the BBL protein domain has been adduced as evidence for barrier-free downhill folding. Other members of the peripheral subunit domain family fold cooperatively at pH 7. To search for possible causes of anomalies in BBL's denaturation behavior, we measured the pH titration of individual residues by heteronuclear NMR. At 298 K, the pKa of His142 was close to that of free histidine at 6.47 ± 0.04, while that of the more buried His166 was highly perturbed at 5.39 ± 0.02. Protonation of His166 is thus energetically unfavorable and destabilizes the protein by ∼ 1.5 kcal/mol. Changes in Cα secondary shifts at pH 5.3 showed a decrease in helicity of the C-terminus of helix 2, where His166 is located, which was accompanied by a measured decrease of 1.1 ± 0.2 kcal/mol in stability from pH 7 to 5.3. Protonation of His166 perturbs, therefore, the structure of BBL. Only ∼ 1% of the structurally perturbed state will be present at the biologically relevant pH 7.6. Experiments at pH 5.3 report on a near-equal mixture of the two different native states. Further, at this pH, small changes of pH and pKa induced by changes in temperature will have near-maximal effects on pH-dependent conformational equilibria and on propagation of experimental error. Accordingly, conventional barrier-limited folding predicts some dispersion of measured thermal unfolding curves of individual residues at pH 5.3.  相似文献   

4.
5.
Expression in Escherichia coli of his-tagged human mevalonate diphosphate decarboxylase (hMDD) has expedited enzyme isolation, characterization, functional investigation of the mevalonate diphosphate binding site, and crystal structure determination (2.4 Å resolution). hMDD exhibits Vmax = 6.1 ± 0.5 U/mg; Km for ATP is 0.69 ± 0.07 mM and Km for (R,S) mevalonate diphosphate is 28.9 ± 3.3 μM. Conserved polar residues predicted to be in the hMDD active site were mutated to test functional importance. R161Q exhibits a ∼1000-fold diminution in specific activity, while binding the fluorescent substrate analog, TNP-ATP, comparably to wild-type enzyme. Diphosphoglycolyl proline (Ki = 2.3 ± 0.3 uM) and 6-fluoromevalonate 5-diphosphate (Ki = 62 ± 5 nM) are competitive inhibitors with respect to mevalonate diphosphate. N17A exhibits a Vmax = 0.25 ± 0.02 U/mg and a 15-fold inflation in Km for mevalonate diphosphate. N17A’s Ki values for diphosphoglycolyl proline and fluoromevalonate diphosphate are inflated (>70-fold and 40-fold, respectively) in comparison with wild-type enzyme. hMDD structure indicates the proximity (2.8 Å) between R161 and N17, which are located in an interior pocket of the active site cleft. The data suggest the functional importance of R161 and N17 in the binding and orientation of mevalonate diphosphate.  相似文献   

6.
Protein aggregates are usually formed by interactions between unfolded or partially unfolded species, and often occur when a protein is denatured by, for example, heat or low pH. In earlier work, we used a Darwinian selection strategy to create human antibody variable domains that resisted heat aggregation. The repertoires of domains were displayed on filamentous phage and denatured (at 80 °C in pH 7.4), and folded domains were selected by binding to a generic ligand after cooling. This process appeared to select for domains with denatured states that resisted aggregation, but the domains only had low free energies of folding (ΔGN-Do = 15-20 kJ/mol at 25 °C in pH 7.4). Here, using the same phage repertoire, we have extended the method to the selection of domains resistant to acid aggregation. In this case, however, the thermodynamic stabilities of selected domains were higher than those selected by thermal denaturation (under both neutral and acidic conditions; ΔGN-Do = 26-47 kJ/mol at 25 °C in pH 7.4, or ΔGN-Do = 27-34 kJ/mol in pH 3.2). Furthermore, we identified a key determinant (Arg28) that increased the aggregation resistance of the denatured states of the domains at low pH without compromising their thermodynamic stabilities. Thus, the selection process yielded domains that combined thermodynamic stability and aggregation-resistant unfolded states. We suggest that changes to these properties are controlled by the extent to which the folding equilibrium is displaced during the process of selection.  相似文献   

7.
The pH dependence of the Fe(III) reduction potential, E0′, for yeast cytochrome c peroxidase (yCcP) and three distal pocket mutants, CcP(H52L), CcP(H52Q), and CcP(R48L/W51L/H52L), has been determined between pH 4 and 8. E0′ values at pH 7.0 for the yCcP, CcP(H52L), CcP(H52Q), and CcP(R48L/W51L/H52L) are − 189, − 170, − 224, and − 146 mV, respectively. A heme-linked ionization in the reduced enzyme affects the reduction potential for yCcP and all three mutants. Apparent pKA values for the heme-linked ionization are 7.5 ± 0.2, 6.5 ± 0.3, 6.4 ± 0.2, and 7.0 ± 0.3 for yCcP and the H52L, H52Q, and R48L/W51L/H52L mutants, respectively. A cooperative, two-proton ionization causing a spectroscopically-detectable transition was observed in the ferrous states of yCcP, CcP(H52L) and CcP(H52Q), with apparent pKA values of 7.7 ± 0.2, 7.4 ± 0.1 and 7.8 ± 0.1, respectively. These data indicate that: (1) the distal histidine in CcP is not the site of proton binding upon reduction of the ferric CcP, (2) the distal histidine is not one of the two groups involved in the cooperative, two-proton ionization observed in ferrous CcP, and (3) the proton-binding site is not involved in the cooperative, two-proton ionization observed in the reduced enzyme.  相似文献   

8.
Galectins are a family of lectins with a conserved carbohydrate recognition domain that interacts with β-galactosides. By binding cell surface glycoconjugates, galectin-1 (gal-1) is involved in cell adhesion and migration processes and is an important regulator of tumor angiogenesis. Here, we used heteronuclear NMR spectroscopy and molecular modeling to investigate lactose binding to gal-1 and to derive solution NMR structures of gal-1 in the lactose-bound and unbound states. Structure analysis shows that the β-strands and loops around the lactose binding site, which are more open and dynamic in the unbound state, fold in around the bound lactose molecule, dampening internal motions at that site and increasing motions elsewhere throughout the protein to contribute entropically to the binding free energy. CD data support the view of an overall more open structure in the lactose-bound state. Analysis of heteronuclear single quantum coherence titration binding data indicates that lactose binds the two carbohydrate recognition domains of the gal-1 dimer with negative cooperativity, in that the first lactose molecule binds more strongly (K1 = 21 ± 6 × 103 M− 1) than the second (K2 = 4 ± 2 × 103 M− 1). Isothermal calorimetry data fit using a sequential binding model present a similar picture, yielding K1 = 20 ± 10 × 103 M− 1 and K2 = 1.67 ± 0.07 × 103 M− 1. Molecular dynamics simulations provide insight into structural dynamics of the half-loaded lactose state and, together with NMR data, suggest that lactose binding at one site transmits a signal through the β-sandwich and loops to the second binding site. Overall, our results provide new insight into gal-1 structure-function relationships and to protein-carbohydrate interactions in general.  相似文献   

9.
Hybrid density functional theory (DFT) and time-dependent DFT (TD-DFT) calculations have been carried out for ozone-water clusters O3(H2O)n (n = 1-4) in order to obtain hydration effects on the absorption spectrum of ozone. The first water molecule in n = 1 is bound to the ozone molecule by an oxygen orientation form in which the oxygen atom of H2O orients the central oxygen atom of O3. In n = 2, the water dimer is bound to O3 and then the cyclic structure is formed as the most stable structure. For n = 3 (or n = 4), the cyclic water trimer (or tetramer) is bound by a hydrogen bond to the ozone molecule. The TD-DFT calculations of O3(H2O)n (n = 0-4) show that the first and second excitation energies of O3 are blue-shifted by the interaction with the water clusters. The magnitude of the spectral shift is largest in n = 2, and the shifts of the excitation energies are +0.07 eV for S1 and +0.13 eV for S2 states. In addition to the spectral shifts (S1 and S2 states), it is suggested that a charge-transfer band is appeared as a low-lying excited state above the S1 and S2 states. The origin of the spectrum shifts was discussed on the basis of theoretical results.  相似文献   

10.
We investigate the hydration state of horse-heart cytochrome c (hh cyt c) in the unfolding process induced by trifluoroacetic acid (TFA). The conformation of hh cyt c changes from the native (N) state (2.9 < pH < 6.0) to the acid-unfolded (UA) state (1.7 < pH < 2.0) to the acid-induced molten globule (A) state (pH ∼1.2). Hydration properties of hh cyt c during this process are measured at 20°C by high-resolution dielectric relaxation (DR) spectroscopy, UV-vis absorbance, and circular dichroism spectroscopy. Constrained water of hh cyt c is observed at every pH as an ∼5-GHz Debye component (DC) (DR time, τD ∼30 ps) and its DR amplitude (DRA) is increased by 77% upon N-to-UA transition, when pH changes from 6.0 to 2.0. Even in the N state, the DRA of the constrained-water component is found to be increased by 22% with decreasing pH from 6.0 to 2.9, suggesting an increase in the accessible surface area of native hh cyt c. Moreover, hypermobile water around native hh cyt c is detected at pH 6.0 as a 19-GHz DC (τD ∼ 8.4 ps < τDW = 9.4 ps), but is not found at other pH values. The DRA signal of constrained water is found to return to the pH 2.9 (N-state) level upon UA-to-A transition. Fast-response water (slightly slower than bulk) around A-state hh cyt c is detected at pH 1.2, and this suggests some accumulation of TFA ions around the peptide chain. Thus, this high-resolution DR spectroscopy study reveals that hh cyt c exhibits significant hydration-state change in the TFA-unfolding process.  相似文献   

11.
A case study on Centaurea gymnocarpa Moris & De Not., a narrow endemic species, was carried out by analyzing its morphological, anatomical, and physiological traits in response to natural habitat stress factors under Mediterranean climate conditions. The results underline that the species is particularly adapted to the environment where it naturally grows. At the plant level, the above-ground/below-ground dry mass (1.73 ± 0.60) shows its investment predominately in the above-ground structure with a resulting total leaf area per plant of 1399 ± 94 cm2. The senescent attached leaves at the base of the plant contribute to limit leaf transpiration by shading soil around the plant. Moreover, the dense C. gymnocarpa leaf pubescence, leaf rolling, the relatively high leaf mass area (LMA = 12.3 ± 1.3 mg cm−2) and leaf tissue density (LTD = 427 ± 44 mg cm−3) contribute to limit leaf transpiration, also postponing leaf death under dry conditions. At the physiological level, a relatively low respiration/photosynthesis ratio (R/PN) in spring results from high R [2.26 ± 0.59 μmol (CO2) m−2 s−1] and PN [12.3 ± 1.5 μmol (CO2) m−2 s−1]. The high photosynthetic nitrogen use efficiency [PNUE = 15.5 ± 0.4 μmol (CO2) g−1 (N) s−1] shows the large amount of nitrogen (N) invested in the photosynthetic machinery of new leaves, associated to a high chlorophyll content (Chl = 35 ± 5 SPAD units). On the contrary, the highest R/PN ratio (1.75 ± 0.19) in summer is due to a significant PN decrease and increase of R in response to drought. The low PNUE [1.5 ± 0.2 μmol (CO2) g−1 (N) s−1] in this season is indicative of a greater N investment in leaf cell walls which may contribute to limit transpiration. On the contrary, the low R/PN ratio (0.05 ± 0.02) in winter is resulting from the limited enzyme activity of the respiratory apparatus [R = 0.23 ± 0.08 μmol (CO2) m−2 s−1] while the low PNUE [3.5 ± 0.2 μmol (CO2) g−1 (N) s−1] suggests that low temperatures additionally limit plant production. The experiment of the imposed water stress confirms that the C. gymnocarpa growth capability is in conformity with the severe conditions of its natural habitat, likewise as it may be the case with others narrow endemic species that have occupied niches with similar extreme conditions.  相似文献   

12.
Escherichiacoli RecBCD is a bipolar DNA helicase possessing two motor subunits (RecB, a 3′-to-5′ translocase, and RecD, a 5′-to-3′ translocase) that is involved in the major pathway of recombinational repair. Previous studies indicated that the minimal kinetic mechanism needed to describe the ATP-dependent unwinding of blunt-ended DNA by RecBCD in vitro is a sequential n-step mechanism with two to three additional kinetic steps prior to initiating DNA unwinding. Since RecBCD can “melt out” ∼ 6 bp upon binding to the end of a blunt-ended DNA duplex in a Mg2+-dependent but ATP-independent reaction, we investigated the effects of noncomplementary single-stranded (ss) DNA tails [3′-(dT)6 and 5′-(dT)6 or 5′-(dT)10] on the mechanism of RecBCD and RecBC unwinding of duplex DNA using rapid kinetic methods. As with blunt-ended DNA, RecBCD unwinding of DNA possessing 3′-(dT)6 and 5′-(dT)6 noncomplementary ssDNA tails is well described by a sequential n-step mechanism with the same unwinding rate (mkU = 774 ± 16 bp s− 1) and kinetic step size (m = 3.3 ± 1.3 bp), yet two to three additional kinetic steps are still required prior to initiation of DNA unwinding (kC = 45 ± 2 s− 1). However, when the noncomplementary 5′ ssDNA tail is extended to 10 nt [5′-(dT)10 and 3′-(dT)6], the DNA end structure for which RecBCD displays optimal binding affinity, the additional kinetic steps are no longer needed, although a slightly slower unwinding rate (mkU = 538 ± 24 bp s− 1) is observed with a similar kinetic step size (m = 3.9 ± 0.5 bp). The RecBC DNA helicase (without the RecD subunit) does not initiate unwinding efficiently from a blunt DNA end. However, RecBC does initiate well from a DNA end possessing noncomplementary twin 5′-(dT)6 and 3′-(dT)6 tails, and unwinding can be described by a simple uniform n-step sequential scheme, without the need for the additional kC initiation steps, with a similar kinetic step size (m = 4.4 ± 1.7 bp) and unwinding rate (mkobs = 396 ± 15 bp s− 1). These results suggest that the additional kinetic steps with rate constant kC required for RecBCD to initiate unwinding of blunt-ended and twin (dT)6-tailed DNA reflect processes needed to engage the RecD motor with the 5′ ssDNA.  相似文献   

13.
We describe a simple method for the determination of heme protein reduction potentials. We use the method to determine the reduction potentials for the PAS-A domains of the regulatory heme proteins human NPAS2 (Em = −115 mV ± 2 mV, pH 7.0) and human CLOCK (Em = −111 mV ± 2 mV, pH 7.0). We suggest that the method can be easily and routinely applied to the determination of reduction potentials across the family of heme proteins.  相似文献   

14.
A model mineralizing system was subjected to magnetic resonance microscopy to investigate how water proton transverse (T2) relaxation times and magnetization transfer ratios can be applied to monitor collagen mineralization. In our model system, a collagen sponge was mineralized with polymer-stabilized amorphous calcium carbonate. The lower hydration and water proton T2 values of collagen sponges during the initial mineralization phase were attributed to the replacement of the water within the collagen fibrils by amorphous calcium carbonate. The significant reduction in T2 values by day 6 (p < 0.001) was attributed to the appearance of mineral crystallites, which were also detected by x-ray diffraction and scanning electron microscopy. In the second phase, between days 6 and 13, magnetic resonance microscopy properties appear to plateau as amorphous calcium carbonate droplets began to coalesce within the intrafibrillar space of collagen. In the third phase, after day 15, the amorphous mineral phase crystallized, resulting in a reduction in the absolute intensity of the collagen diffraction pattern. We speculate that magnetization transfer ratio values for collagen sponges, with similar collagen contents, increased from 0.25 ± 0.02 for control strips to a maximum value of 0.31 ± 0.04 at day 15 (p = 0.03) because mineral crystals greatly reduce the mobility of the collagen fibrils.  相似文献   

15.
In vitro antitumour activity of the [Pt(ox)(Ln)2] (1-7) and [Pd(ox)(Ln)2] (8-14) oxalato (ox) complexes involving N6-benzyl-9-isopropyladenine-based N-donor carrier ligands (Ln) against ovarian carcinoma (A2780), cisplatin resistant ovarian carcinoma (A2780cis), malignant melanoma (G-361), lung carcinoma (A549), cervix epitheloid carcinoma (HeLa), breast adenocarcinoma (MCF7) and osteosarcoma (HOS) human cancer cell lines was studied. Some of the tested complexes were even several times more cytotoxic as compared with cisplatin employed as a positive control. The improved cytotoxic effect was demonstrated for the platinum(II) complexes 3 (IC50 = 3.2 ± 1.0 μM and 3.2 ± 0.6 μM) and 5 (IC50 = 4.0 ± 1.0 μM and 4.1 ± 1.4 μM) against A2780 and A2780cis, as compared with 11.5 ± 1.6 μM, and 30.3 ± 6.1 μM determined for cisplatin, respectively. The significant in vitro cytotoxicity against MCF7 (IC50 = 8.2 ± 3.8 μM for 12) and A2780 (IC50 = 5.4 ± 1.2 μM for 14) was evaluated for the palladium(II) oxalato complexes, which again exceeded cisplatin, whose IC50 equalled 19.6 ± 4.3 μM against the MCF7 cells. Selected complexes were also screened for their in vitro cytotoxic effect in primary cultures of human hepatocytes and they were found to be non-hepatotoxic.  相似文献   

16.
Endothelin-1 (ET-1) induces contraction of vascular smooth muscle through binding to endothelin type A receptor (ETAR). COS-7 cells stably expressing high levels of the ETAR were established (designated COS-7(ETAR)). The COS-7(ETAR) cell bound [125I]ET-1 with a Kd of 932 ± 161 pM and a Bmax of 74 ± 13 fmol/2 × 105 cells. [125I]ET-1 binding was inhibited by ET-1 and the ETAR antagonist BQ-610, but not by the endothelin type B receptor (ETBR) antagonist BQ-788. In clones expressing two ETAR mutants containing D46N or R53Q substitutions in the first extracellular domain of the receptor, [125I]ET-1 binding activity was dramatically reduced. This suggests that these single amino acid substitutions alter the three-dimensional structure of the ligand-binding domain of the ETAR. Using COS-7(ETAR) cell, we showed that Ca2+ or Mg2+ was essential for ET-1 binding to the ETAR and that ET-1 treatment induced postreceptor signaling, that is, intracellular accumulation of cyclic AMP (cAMP) and Ca2+ mobilization. The COS-7(ETAR) established in this study will be a useful tool for screening ET-1 antagonists for treating hypertension.  相似文献   

17.
Flavodiiron proteins (FDPs) are enzymes identified in prokaryotes and a few pathogenic protozoa, which protect microorganisms by reducing O2 to H2O and/or NO to N2O. Unlike most prokaryotic FDPs, the protozoan enzymes from the human pathogens Giardia intestinalis and Trichomonas vaginalis are selective towards O2. UV/vis and EPR spectroscopy showed that, differently from the NO-consuming bacterial FDPs, the Giardia FDP contains an FMN with reduction potentials for the formation of the single and the two-electron reduced forms very close to each other (E1 = −66 ± 15 mV and E2 = −83 ± 15 mV), a condition favoring destabilization of the semiquinone radical. Giardia FDP contains also a non-heme diiron site with significantly up-shifted reduction potentials (E1 = +163 ± 20 mV and E2 = +2 ± 20 mV). These properties are common to the Trichomonas hydrogenosomal FDP, and likely reflect yet undetermined subtle structural differences in the protozoan FDPs, accounting for their marked O2 specificity.  相似文献   

18.
The reaction of Cu(ClO4)2 · 6H2O with bis(3-aminopropyl)methylamine and sodium dicyanamide in aqueous medium results in the formation of a dimeric dicyanamide complex of Cu(II), [Cu2(medpt)2(dca)2](ClO4)2. The single crystal X-ray structure reveals that the dinuclear entities are extended to form a supramolecular 1D ladder by H-bonding. Each dinuclear entity is joined to the adjacent unit via the perchlorate anion. Variable temperature magnetic study was performed and the best-fit parameters are J = −1.20 ± 0.02 cm−1, g = 2.08 ± 0.01 with R = 2 × 10−5. These clearly indicate the antiferromagnetic interaction between the Cu(II) center.  相似文献   

19.
X-ray diffuse scattering was measured from oriented stacks and unilamellar vesicles of dioleoylphosphatidylcholine lipid bilayers to obtain the temperature dependence of the structure and of the material properties. The area/molecule, A, was 75.5 Å2 at 45°C, 72.4 Å2 at 30°C, and 69.1 Å2 at 15°C, which gives the area expansivity αA = 0.0029/deg at 30°C, and we show that this value is in excellent agreement with the polymer brush theory. The bilayer becomes thinner with increasing temperature; the contractivity of the hydrocarbon portion was αDc = 0.0019/deg; the difference between αA and αDc is consistent with the previously measured volume expansivity αVc = 0.0010/deg. The bending modulus KC decreased as exp(455/T) with increasing T (K). Our area compressibility modulus KA decreased with increasing temperature by 5%, the same as the surface tension of dodecane/water, in agreement again with the polymer brush theory. Regarding interactions between bilayers, the compression modulus B as a function of interbilayer water spacing DW was found to be nearly independent of temperature. The repulsive fluctuation pressure calculated from B and KC increased with temperature, and the Hamaker parameter for the van der Waals interaction was nearly independent of temperature; this explains why the fully hydrated water spacing, DW, that we obtain from our structural results increases with temperature.  相似文献   

20.
pH-dependent (pH 6.0–8.0) quaternary structural changes of ferric Vitreoscilla hemoglobin (VHb) have been investigated using dynamic light scattering. The VHb exhibits a monomeric state under neutral conditions at pH 7.0, while the protein forms distinct homodimeric species at pH 6.0 and 8.0, respectively. The dissociation constant obtained using the Bio-Layer Interferometry technology indicates that, at pH 7.0, the monomer–monomer dissociation of VHb is about 6-fold or 5-fold higher (KD = 6.34 μM) compared with that at slightly acidic pH (KD = 1.05 μM) or slightly alkaline pH (KD = 1.22 μM). The pH-dependent absorption spectra demonstrate that the heme microenvironment of VHb is sensitive to the changes of pH value. The maximum absorption band of heme group of VHb shifts from 402 nm to 407 nm when pH changes from 6.0 to 8.0. In addition, the fluorescence emission spectra of VHb, taken at excitation wavelength of 295 nm, suggest that the single Trp122 fluorescence quantum yields in VHb are decreased due to the formation of the homodimeric species. However, the circular dichroism spectra data display that the secondary structures of VHb are little affected by pH transitions. The pH-dependent peroxidase activity of VHb was also investigated in this study. The optimum pH for VHb using 2,2′-azino-bis-(3-ethylbenzothiazoline-6-sulfonic acid) as substrate is 7.0, which implies that the monomer state of VHb would exhibit better peroxidase activity than the homodimeric species of VHb at pH 6.0 and 8.0.  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