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.
The dissociation kinetics of the europium(III) complex with H8dotp ligand was studied by means of molecular absorption spectroscopy in UV region at ionic strength 3.0 mol dm−3 (Na,H)ClO4 and in temperature region 25-60 °C. Time-resolved laser-induced fluorescence spectroscopy (TRLIFS) was employed in order to determine the number of water molecules in the first coordination sphere of the europium(III) reaction intermediates and the final products. This technique was also utilized to deduce the composition of reaction intermediates in course of dissociation reaction simultaneously with calculation of rate constants and it demonstrates the elucidation of intimate reaction mechanism. The thermodynamic parameters for the formation of kinetic intermediate (ΔH0 = 11 ± 3 kJ mol−1, ΔS0 = 41 ± 11 J K−1 mol−1) and the activation parameters (Ea = 69 ± 8 kJ mol−1, ΔH = 67 ± 8 kJ mol−1, ΔS = −83 ± 24 J K−1 mol−1) for the rate-determining step describing the complex dissociation were determined. The mechanism of proton-assisted reaction was proposed on the basis of the experimental data.  相似文献   

2.
Kinetic studies of X exchange on [AuX4] square-planar complexes (where X=Cl and CN) were performed at acidic pH in the case of chloride system and as a function of pH for the cyanide one. Chloride NMR study (330-365 K) gives a second-order rate law on [AuCl4] with the kinetic parameters: (k2Au,Cl)298=0.56±0.03 s−1 mol−1 kg; ΔH2‡ Au,Cl=65.1±1 kJ mol−1; ΔS2‡ Au,Cl=−31.3±3 J mol−1 K−1 and ΔV2 Au,Cl=−14±2 cm3 mol−1. The variable pressure data clearly indicate the operation of an Ia or A mechanism for this exchange pathway. The proton exchange on HCN was determined by 13C NMR as a function of pH and the rate constant of the three reaction pathways involving H2O, OH and CN were determined: k0HCN,H=113±17 s−1, k1HCN,H=(2.9±0.7)×109 s−1 mol−1 kg and k2HCN,H=(0.6±0.2)×106 s−1 mol−1 kg at 298.1 K. The rate law of the cyanide exchange on [Au(CN)4] was found to be second order with the following kinetic parameters: (k2Au,CN)298=6240±85 s−1 mol−1 kg, ΔH2 Au,CN=40.0±0.8 kJ mol−1, ΔS2 Au,CN=−37.8±3 J mol−1 K−1 and ΔV2 Au,CN=+2±1 cm3 mol−1. The rate constant observed varies about nine orders of magnitude depending on the pH and HCN does not act as a nucleophile. The observed rate constant of X exchange on [AuX4] are two or three orders of magnitude faster than the Pt(II) analogue.  相似文献   

3.
The reaction of [(η7-C7H7)Zr(η5-C5H5)] with two Lewis bases, tetramethylimidazolin-2-ylidene and PMe3, is reported and their stability probed via spectroscopic and theoretical methods. The strongly σ-basic N-heterocyclic carbene forms a stable adduct which has been structurally characterised, whilst the PMe3 ligand coordinates weakly to the metal centre. Variable temperature 31P NMR spectroscopy has been used to determine the activation energy for this process (ΔG = 40.5 ± 1.9 kJ mol−1). DFT calculations have been performed on both complexes and the structures discussed. In addition, the enthalpies for the formation of these compounds have been calculated [ΔH0(Zr-IMe) = −56.3 kJ mol−1; ΔH0(Zr-PMe3) = −2.3 kJ mol−1] and show that the N-heterocyclic carbene forms a thermodynamically much more stable adduct than that with PMe3.  相似文献   

4.
Reaction of the five-coordinate trigonal-bipyramidal platinum(II) complex, [Pt(pt)(pp3)](BF4) (pt = 1-propanethiolate, pp3 = tris[2-(diphenylphosphino)ethyl]phosphine), with I in chloroform gave the five-coordinate square-pyramidal complex with a dissociated terminal phosphino group and an apically coordinated iodide ion in equilibrium. The thermodynamic parameters for the equilibrium between the trigonal-bipyramidal and square-pyramidal geometries, [Pt(pt)(pp3)]+ + I ? [PtI(pt) (pp3)], and the kinetic parameters for the chemical exchange were obtained as follows: , ΔH0 = − 10 ± 2.4 kJ mol−1, ΔS0 = − 36 ± 10 J K−1 mol−1, , ΔH = 34 ± 4.7 kJ mol−1, ΔS = − 50 ± 21 J K−1 mol−1. The square-planar trinuclear platinum(II) complex was formed by bridging reaction of one of the terminal phosphino groups of trigonal-bipyramidal [PtCl(pp3)]Cl with trans-[PtCl2(NCC6H5)2] in chloroform. From these facts, ligand substitution reactions of [PtX(pp3)]+ (X = monodentate anion) are expected to proceed via an intermediate with a dissociated phosphino group. The rate constants for the chloro-ligand substitution reactions of [PtCl(pp3)]+ with Br and I in chloroform approached the respective limiting values as concentrations of the entering halide ions are increased. These kinetic results confirmed the preassociation mechanism in which the square pyramidal intermediate with a dissociated phosphino group and an apically coordinated halide ion is present in the rapid pre-equilibrium.  相似文献   

5.
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   

6.
In order to examine the effects of coordinated hydroxide ion and free hydroxide ion in configurational conversion of a tetraamine macrocyclic ligand complex, the kinetics of the cis-to-planar interconversion of cis-[Ni(isocyclam)(H2O)2]2+ (isocyclam, 1,4,7,11-tetraazacyclotetradecane) has been studied spectrophotometrically in basic aqueous solution. The interconversion requires the inversion of one sec-NH center of the folded cis-complex to have the planar species. Kinetic data are satisfactorily fitted by the rate law, R = kOH[OH][cis-[Ni(isocyclam)(H2O)2]2+], where kOH = 3.84 × 103 dm3 mol−1 s−1 at 25.0 ± 0.1 °C with I = 0.10 mol dm−3 (NaClO4). The large ΔH, 61.7 ± 3.2 kJ mol−1, and the large positive ΔS, 30.2 ± 10.8 J K−1 mol−1, strongly support a free-base-catalyzed mechanism for the reaction.  相似文献   

7.
α-Amylase from Sorghum bicolor, is reversibly unfolded by chemical denaturants at pH 7.0 in 50 mM Hepes containing 13.6 mM calcium and 15 mM DTT. The isothermal equilibrium unfolding at 27 °C is characterized by two state transition with ΔG (H2O) of 16.5 kJ mol−1 and 22 kJ mol−1, respectively, at pH 4.8 and pH 7.0 for GuHCl and ΔG (H2O) of 25.2 kJ mol−1 at pH 4.8 for urea. The conformational stability indicators such as the change in excess heat capacity (ΔCp), the unfolding enthalpy (Hg) and the temperature at ΔG = 0 (Tg) are 17.9 ± 0.7 kJ mol−1 K−1, 501.2 ± 18.2 kJ mol1 and 337.3 ± 6.9 K at pH 4.8 and 14.3 ± 0.5 kJ mol−1 K−1, 509.3 ± 21.7 kJ mol−1 and 345.4 ± 4.8 K at pH 7.0, respectively. The reactivity of the conserved cysteine residues, during unfolding, indicates that unfolding starts from the ‘B’ domain of the enzyme. The oxidation of cysteine residues, during unfolding, can be prevented by the addition of DTT. The conserved cysteine residues are essential for enzyme activity but not for the secondary and tertiary fold acquired during refolding of the denatured enzyme. The pH dependent stability described by ΔG (H2O) and the effect of salt on urea induced unfolding confirm the role of electrostatic interactions in enzyme stability.  相似文献   

8.
In N,N-dimethylformamide (DMF), 1,4,7-tris((S)-2-hydroxy-3-phenylpropyl)-1,4,7-triazacyclononane forms metal complexes, [M(S-thppc9)]+, for which log K (dm3 mol−1)=3.01, 2.65, 2.66, 2.65, 2.42 and 7.59 (all±0.05) where M+=Li+, Na+, K+, Rb+, Cs+ and Ag+, respectively. Variable temperature 13C{1H} NMR spectroscopy shows that the interchange between equivalent forms of a single diastereomer occurs for [Li(S-thppc9)]+ and [Na(S-thppc9)]+ characterised by: k=43±5 and 2900±100 s−1, at 298.2 K, ΔH=22.5±1.6 and 33.8±1.6 kJ mol−1, and ΔS=−133±5 and −59±6 J K−1 mol−1, respectively. Gas phase ab initio modelling shows these complexes and their K+ analogue to preferentially form distorted trigonal prismatic Λ, Δ, and Λ diastereomers, respectively.  相似文献   

9.
The kinetic results of the oxidative addition of iodomethane to Bu4N[Ir2(μ-Dcbp)(cod)2] (Dcbp = 3,5-dicarboxylatepyrazolate anion) show that oxidative addition can occur via a direct equilibrium pathway (K1 = 88(22) acetone, 51(3) 1,2-dichloroethane, 55(4) dichloromethane, 52(12) acetonitrile and 43(5) M−1 chloroform) or a solvent-assisted pathway (k2, k3). Oxidative addition occurs mainly along the direct pathway, which is a factor 10-40 faster than the solvent-assisted pathway. The observed solvent effect cannot be attributed to the donosity or polarity of the solvents. The fairly negative ΔS value (−110(7) J K−1 mol−1) and the positive ΔH value (+47(2) kJ mol−1) for the oxidative addition step are indicative of an associative process.  相似文献   

10.
Two isomers of the N,O-coordinated acetylpyrrolyl complex [Ru(PPh3)2(CO)(NC4H3C(O)CH3)H] {cis-N,H (1) and trans-N,H (2)} have been prepared as models for catalytic intermediates in the Murai reaction. Complex 2 isomerises to 1 upon heating via a dissociative pathway (ΔH = 195 ± 41 kJ mol−1; ΔS = 232 ± 62 J mol−1 K−1); the mechanism of this process has been modeled using density functional calculations. Complex 2 displays moderate catalytic activity for the Murai coupling of 2′-methylacetophenone with trimethylvinylsilane, but 1 proved to be catalytically inactive under the same conditions.  相似文献   

11.
To characterize driving forces and driven processes in formation of a large-interface, wrapped protein-DNA complex analogous to the nucleosome, we have investigated the thermodynamics of binding the 34-base pair (bp) H′ DNA sequence to the Escherichia coli DNA-remodeling protein integration host factor (IHF). Isothermal titration calorimetry and fluorescence resonance energy transfer are applied to determine effects of salt concentration [KCl, KF, K glutamate (KGlu)] and of the excluded solute glycine betaine (GB) on the binding thermodynamics at 20 °C. Both the binding constant Kobs and enthalpy ΔH°obs depend strongly on [salt] and anion identity. Formation of the wrapped complex is enthalpy driven, especially at low [salt] (e.g., ΔHoobs = − 20.2 kcal·mol− 1 in 0.04 M KCl). ΔH°obs increases linearly with [salt] with a slope (dΔH°obs/d[salt]), which is much larger in KCl (38 ± 3 kcal·mol− 1 M− 1) than in KF or KGlu (11 ± 2 kcal·mol− 1 M− 1). At 0.33 M [salt], Kobs is approximately 30-fold larger in KGlu or KF than in KCl, and the [salt] derivative SKobs = dlnKobs/dln[salt] is almost twice as large in magnitude in KCl (− 8.8 ± 0.7) as in KF or KGlu (− 4.7 ± 0.6).A novel analysis of the large effects of anion identity on Kobs, SKobs and on ΔH°obs dissects coulombic, Hofmeister, and osmotic contributions to these quantities. This analysis attributes anion-specific differences in Kobs, SKobs, and ΔH°obs to (i) displacement of a large number of water molecules of hydration [estimated to be 1.0(± 0.2) × 103] from the 5340 Å2 of IHF and H′ DNA surface buried in complex formation, and (ii) significant local exclusion of F and Glu from this hydration water, relative to the situation with Cl, which we propose is randomly distributed. To quantify net water release from anionic surface (22% of the surface buried in complexation, mostly from DNA phosphates), we determined the stabilizing effect of GB on Kobs: dlnKobs/d[GB]  = 2.7 ± 0.4 at constant KCl activity, indicating the net release of ca. 150 H2O molecules from anionic surface.  相似文献   

12.
The X-ray crystal structures of two related trans-N2S2 copper macrocycles are reported. One was isolated with the copper in the divalent form and the other with copper in its univalent form affording a valuable insight into the changes of geometry and metrical parameters that occur during redox processes in macrocyclic copper complexes. A variable temperature NMR study of the copper(I) complex is reported, indicative of a chair-boat conformational change within the alkyl chain backbone of the macrocycle. It was possible to extract the relevant kinetic and thermodynamic parameters (ΔG, 57.8 kJ mol−1; ΔH, 52.1 kJ mol−1; ΔS, −19.2 J K−1 mol−1) for this process at 298 K. DFT molecular orbital calculations were used to confirm these observations and to calculate the energy difference (26.2 kJmol−1) between the copper(I) macrocycle in a planar and a distorted tetrahedral disposition.  相似文献   

13.
The kinetics of the complexation of Ni(II) with 1,10-phenanthroline(phen), 4,7-dimethyl-1,10-phenanthroline(dmphen), and 5-nitro-1,10-phenanthroline(NO2phen) in acetonitrile-water mixed solvents of acetonitrile mole fraction xAN = 0, 0.05, 0.1, 0.2 and 0.3 at 288, 293, 298 and 303 K have been studied by stopped-flow method at ionic strength of 1.0 (NaClO4) and pH 7.4. The corresponding activation enthalpy, entropy, and free energy were determined from the observed rate constants. The complexation of Ni(II) with the three ligands has comparable observed rate constants; in pure water the observed rate constants are (×103 dm3 mol−1 s−1) 2.31, 2.57, and 1.38 for phen, dmphen and NO2phen, respectively. The corresponding activation parameters for the three ligands are, however, considerably different; in pure water the ΔHS (kJ mol−1/J K−1 mol−1) are 44.7/−30.2, 19.5/−114.1, and 32.2/−76.9 for phen, dmphen, and NO2phen, respectively. The effects of solvent composition on the kinetics are also markedly different for the three ligands. The ΔH and ΔS showed a minimum at xAN = 0.1 for phen; for dmphen and NO2phen, however, maxima at xAN = 0.2 were observed. Nevertheless, there is an effective enthalpy-entropy compensation for the ΔHS of all the three ligands, demonstrating the significant effects of the changes in solvation and solvent structure on the complexation kinetics. As the rate-determining step of Ni(II) complexation is the dissociation of a water molecule from Ni(II), the solvent and ligand dependencies in the Ni(II) complexation kinetics are ascribed to the change in solvation status of the ligands and the altered solvent structures upon changing solvent composition.  相似文献   

14.
Homoleptic eight- and nine-coordinate U(IV) perchlorate complexes with sulfoxide ligands have been characterized crystallographically. Crystals of [U(dmso)8](ClO4)4 · 0.75CH3NO2, [U(dmso)9](ClO4)4 · 4dmso (dmso = dimethyl sulfoxide), and [U(tmso)8](ClO4)4 · 2tmso (tmso = tetramethylene sulfoxide) were found to have dodecahedral, tricapped trigonal prismatic, and square antiprismatic geometries, respectively. Average U-O bond distances in [U(dmso)8](ClO4)4 · 0.75CH3NO2, [U(dmso)9](ClO4)4 · 4dmso, and [U(tmso)8](ClO4)4 · 2tmso are 2.35(3), 2.41 (4), and 2.35(3) Å, respectively. Furthermore, it was found that [U(dmso)8]4+ is in equilibrium with [U(dmso)9]4+ in CH3NO2 solution containing dmso. Thermodynamic parameters for such an equilibrium are as follows: K (25 °C) = 3.4 ± 0.2 dm3 mol−1, ΔH = −54.9 ± 4.5 kJ mol−1, and ΔS = −174 ± 15 J K−1 mol−1.  相似文献   

15.
Human serum albumin (HSA) is a monomeric allosteric protein. Here, the effect of ibuprofen on denitrosylation kinetics (koff) and spectroscopic properties of HSA-heme-Fe(II)-NO is reported. The koff value increases from (1.4 ± 0.2) × 10−4 s−1, in the absence of the drug, to (9.5 ± 1.2) × 10−3 s−1, in the presence of 1.0 × 10−2 M ibuprofen, at pH 7.0 and 10.0 °C. From the dependence of koff on the drug concentration, values of the dissociation equilibrium constants for ibuprofen binding to HSA-heme-Fe(II)-NO (K1 = (3.1 ± 0.4) × 10−7 M, K2 = (1.7 ± 0.2) × 10−4 M, and K3 = (2.2 ± 0.2) × 10−3 M) were determined. The K3 value corresponds to the value of the dissociation equilibrium constant for ibuprofen binding to HSA-heme-Fe(II)-NO determined by monitoring drug-dependent absorbance spectroscopic changes (H = (2.6 ± 0.3) × 10−3 M). Present data indicate that ibuprofen binds to the FA3-FA4 cleft (Sudlow’s site II), to the FA6 site, and possibly to the FA2 pocket, inducing the hexa-coordination of HSA-heme-Fe(II)-NO and triggering the heme-ligand dissociation kinetics.  相似文献   

16.
The interaction between CpRuH(dppe) and a series of proton donors (HA) of increasing strength: CFH2CH2OH (MFE), CF3CH2OH (TFE), (CF3)2CHOH (HFIP), p-nitrophenol, CF3COOH and HBF4 has been investigated spectroscopically by variable-temperature IR, UV-Vis, and NMR spectroscopy in solvents of differing polarity (n-hexane, dichloromethane and their mixture). The low-temperature IR study shows the establishment of a hydrogen-bond which involves the hydride ligand as the proton accepting site. The basicity factor Ej for the hydride was found to be 1.39. All techniques indicate that an equilibrium exists between the dihydrogen-bonded complex and the cationic dihydrogen complex, [CpRu(η2-H2)(dppe)]+, the formation of which is shown here for the first time. The proton transfer from HFIP is characterized by ΔH° = −8.1 ± 0.6 kcal mol−1 and ΔS° = −17 ± 3 eu. The activation parameters for the subsequent irreversible isomerization leading to the classical dihydride complex, [CpRu(H)2(dppe)]+, are ΔH = 20.9 ± 0.8 kcal mol−1 and ΔS = 9 ± 3 eu as determined from 1H NMR spectroscopy for protonation by HBF4. Computational results at the DFT/B3PW91 level confirm the experimentally observed hydride basicity increase on descending the Group from iron to ruthenium and also the formation of the non-classical complex as an intermediate, prior to the thermodynamically favored dihydride.  相似文献   

17.
Substitution reaction of fac-[FeII(CN)2(CO)3I] with triphenylphosphine (PPh3) produced mono phosphine substituted complex cis-cis-[FeII(CN)2(CO)2(PPh3)I]. Crystal structure of the product showed that carbonyl positioned trans- to iodide was replaced by PPh3. The substitution reaction was monitored by quantitative infrared spectroscopic method, and the rate law for the substitution reaction was determined to be rate = k[[FeII(CN)2(CO)2(PPh3)I]][PPh3]. Transition state enthalpy and entropy changes were obtained from Eyring equation k = (kBT/h)exp(−ΔH/RT + ΔS/R) with ΔH = 119(4) kJ mol−1 and ΔS = 102(10) J mol−1 K−1. Positive transition state entropy change suggests that the substitution reaction went through a dissociative pathway.  相似文献   

18.
Despite the widespread presence of the globin fold in most living organisms, only eukaryotic globins have been employed as model proteins in folding/stability studies so far. This work introduces the first thermodynamic and kinetic characterization of a prokaryotic globin, that is, the apo form of the heme-binding domain of flavohemoglobin (apoHmpH) from Escherichia coli. This bacterial globin has a widely different sequence but nearly identical structure to its eukaryotic analogues. We show that apoHmpH is a well-folded monomeric protein with moderate stability at room temperature [apparent ΔG°UN(w) = − 3.1 ± 0.3 kcal mol− 1; mUN = − 1.7 kcal mol− 1 M− 1] and predominant α-helical structure. Remarkably, apoHmpH is the fastest-folding globin known to date, as it refolds about 4- to 16-fold more rapidly than its eukaryotic analogues (e.g., sperm whale apomyoglobin and soybean apoleghemoglobin), populating a compact kinetic intermediate (βI = 0.9 ± 0.2) with significant helical content. Additionally, the single Trp120 (located in the native H helix) becomes locked into a fully native-like environment within 6 ms, suggesting that this residue and its closest spatial neighbors complete their folding at ultrafast (submillisecond) speed. In summary, apoHmpH is a bacterial globin that shares the general folding scheme (i.e., a rapid burst phase followed by slower rate-determining phases) of its eukaryotic analogues but displays an overall faster folding and a kinetic intermediate with some fully native-like traits. This study supports the view that the general folding features of bacterial and eukaryotic globins are preserved through evolution while kinetic details differ.  相似文献   

19.
A new ligand, N,N′-dibenzylethane-1,2-diamine (L) and its four transition metal(II) complexes, ML2(OAc)2 · 2H2O (M = Cu, Ni, Zn, Co), have been synthesized and characterized by elemental analysis, mass spectra, molar conductivity, NMR and IR. Moreover, the crystals structure of Cu(II) and Ni(II) complexes characterized by single crystal X-ray diffraction showed that the complexes have a similar molecular structure. Ni(II) has an regular octahedral coordination environment complexes, but typical Jahn Teller effect influenced Cu(II) in an elongated octahedral environment. The interaction between complexes and calf thymus DNA were studied by UV and fluorescence spectra measure, which showed that the binding mode of complexes with DNA is intercalation. Under physiological pH condition, the effects of Cu(OAc)2L2 · 2H2O and Ni(OAc)2L2 · 2H2O on human serum albumin were examined by fluorescence. The results of spectroscopic measurements suggested that the hydrophobic interaction is the predominant intermolecular force. The enthalpy change ΔH0 and the entropy change ΔS0 of Cu(OAc)2L2 · 2H2O and Ni(OAc)2L2 · 2H2O were calculated to be −11.533 kJ mol−1 and 46.339 J mol−1 K−1, −11.026 kJ mol−1 and 46.396 J mol−1 K−1, respectively, according to the Scatchard’s equation. The quenching mechanism and the number of binding site (n ≈ 1) were also obtained from fluorescence titration data.  相似文献   

20.
Folding mechanisms and stability of membrane proteins are poorly understood because of the known difficulties in finding experimental conditions under which reversible denaturation could be possible. In this work, we describe the equilibrium unfolding of Archaeoglobus fulgidus CopA, an 804-residue α-helical membrane protein that is involved in transporting Cu+ throughout biological membranes. The incubation of CopA reconstituted in phospholipid/detergent mixed micelles with high concentrations of guanidinium hydrochloride induced a reversible decrease in fluorescence quantum yield, far-UV ellipticity, and loss of ATPase and phosphatase activities. Refolding of CopA from this unfolded state led to recovery of full biological activity and all the structural features of the native enzyme. CopA unfolding showed typical characteristics of a two-state process, with ΔGw° = 12.9 kJ mol 1, = 4.1 kJ mol− 1 M− 1, Cm = 3 M, and ΔCpw° = 0.93 kJ mol− 1 K− 1. These results point out to a fine-tuning mechanism for improving protein stability. Circular dichroism spectroscopic analysis of the unfolded state shows that most of the secondary and tertiary structures were disrupted. The fraction of Trp fluorescence accessible to soluble quenchers shifted from 0.52 in the native state to 0.96 in the unfolded state, with a significant spectral redshift. Also, hydrophobic patches in CopA, mainly located in the transmembrane region, were disrupted as indicated by 1-anilino-naphtalene-8-sulfonate fluorescence. Nevertheless, the unfolded state had a small but detectable amount of residual structure, which might play a key role in both CopA folding and adaptation for working at high temperatures.  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