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 binding of the gelsolin P2 peptide (residues 150-169) with lysophosphatidic acid (LPA) and lipopolysaccharide (LPS) was investigated by isothermal titration calorimetry. P2 binds to LPS with higher affinity than to LPA. For the interaction of 1-oleoyl-LPA with P2 in the absence of salt, K(d) and deltaH degrees were 920 nM and -2.07 kcal/mol, respectively, at pH 7.4 and 25 degrees C. For the interaction of lipopolysaccharide (LPS) from P. aeruginosa with P2 under the same conditions, K(d) was 177 nM and deltaH degrees was -7.6 kcal/mol.  相似文献   

2.
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   

3.

Background

Trypanosoma brucei, responsible for African sleeping sickness, is a lethal parasite against which there is need for new drug protocols. It is therefore relevant to attack possible biomedical targets with specific preparations and since arginine kinase does not occur in humans but is present in the parasite it becomes a suitable target.

Methods

Fluorescence quenching, thermodynamic analysis and FRET have shown that arginine kinase from T. brucei interacted with silver or gold nanoparticles.

Results

The enzyme only had one binding site. At 25 °C the dissociation (Kd) and Stern–Volmer constants (KSV) were 15.2 nM, 0.058 nM− 1 [Ag]; and 43.5 nM, 0.052 nM− 1 [Au] and these decreased to 11.2 nM, 0.041 nM− 1 [Ag]; and 24.2 nM, 0.039 nM− 1 [Au] at 30 °C illustrating static quenching and the formation of a non-fluorescent fluorophore–nanoparticle complex. Silver nanoparticles bound to arginine kinase with greater affinity, enhanced fluorescence quenching and easier access to tryptophan molecules than gold. Negative ΔH and ΔG values implied that the interaction of both Ag and Au nanoparticles with arginine kinase was spontaneous with electrostatic forces. FRET confirmed that the nanoparticles were bound 2.11 nm [Ag] and 2.26 nm [Au] from a single surface tryptophan residue.

Conclusions

The nanoparticles bind close to the arginine substrate through a cysteine residue that controls the electrophilic and nucleophilic characters of the substrate arginine–guanidinium group crucial for enzymatic phosphoryl transfer between ADP and ATP.

General significance

The nanoparticles of silver and gold interact with arginine kinase from T. brucei and may prove to have far reaching consequences in clinical trials.  相似文献   

4.
A new, to our knowledge, group contribution method based on the group contribution method of Mavrovouniotis is introduced for estimating the standard Gibbs free energy of formation (ΔfG′°) and reaction (ΔrG′°) in biochemical systems. Gibbs free energy contribution values were estimated for 74 distinct molecular substructures and 11 interaction factors using multiple linear regression against a training set of 645 reactions and 224 compounds. The standard error for the fitted values was 1.90 kcal/mol. Cross-validation analysis was utilized to determine the accuracy of the methodology in estimating ΔrG′° and ΔfG′° for reactions and compounds not included in the training set, and based on the results of the cross-validation, the standard error involved in these estimations is 2.22 kcal/mol. This group contribution method is demonstrated to be capable of estimating ΔrG′° and ΔfG′° for the majority of the biochemical compounds and reactions found in the iJR904 and iAF1260 genome-scale metabolic models of Escherichia coli and in the Kyoto Encyclopedia of Genes and Genomes and University of Minnesota Biocatalysis and Biodegradation Database. A web-based implementation of this new group contribution method is available free at http://sparta.chem-eng.northwestern.edu/cgi-bin/GCM/WebGCM.cgi.  相似文献   

5.
G-quadruplexes (GQ) are formed by the association of guanine-rich stretches of DNA. Certain small molecules can influence kinetics and thermodynamics of this association. Understanding the mechanism of ligand-assisted GQ folding is necessary for the design of more efficient cancer therapeutics. The oligonucleotide d(TAGGG)2 forms parallel bimolecular GQ in the presence of ≥66 mM K+; GQs are not formed under Na+, Li+ or low K+ conditions. The thermodynamic parameters for GQ folding at 60 μM oligonucleotide and 100 mM KCl are ΔH = −35 ± 2 kcal mol−1 and ΔG310 = −1.4 kcal mol−1. Quadruplex [d(TAGGG)2]2 binds 2-3 K+ ions with Kd of 0.5 ± 0.2 mM. Our work addresses the question of whether metal free 5,10,15,20-tetrakis(N-methyl-4-pyridyl)porphyrin (TMPyP4) and its Zn(II), Cu(II), and Pt(II) derivatives are capable of facilitating GQ folding of d(TAGGG)2 from single stranded, or binding to preformed GQ, using UV-vis and circular dichroism (CD) spectroscopies. ZnTMPyP4 is unique among other porphyrins in its ability to induce GQ structure of d(TAGGG)2, which also requires at least a low amount of potassium. ZnTMPyP4 binds with 2:1 stoichiometry possibly in an end-stacking mode with a ∼106 M−1 binding constant, determined through UV-vis and ITC titrations. This process is entropically driven and has ΔG298 of −8.0 kcal mol−1. TMPyP4 binds with 3:1 stoichiometry and Ka of ∼106 M−1. ZnTMPyP4 and TMPyP4 are efficient stabilizers of [d(TAGGG)2]2 displaying ΔT1/2 of 13.5 and 13.8 °C, respectively, at 1:2 GQ to porphyrin ratio; CuTMPyP4 shows a much weaker effect (ΔT1/2 = 4.7 °C) and PtTMPyP4 is weakly destabilizing (ΔT1/2 = −2.9 °C). The selectivity of ZnTMPyP4 for GQ versus dsDNA is comparable to that of TMPyP4. The ability of ZnTMPyP4 to bind and stabilize GQ, to induce GQ formation, and speed up its folding may suggest an important biological activity for this molecule.  相似文献   

6.
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   

7.
A thermodynamic study of the inclusion process between 2-chlorobenzophenone (2ClBP) and cyclomaltoheptaose (β-cyclodextrin, β-CD) was performed using UV–vis spectroscopy, reversed-phase liquid chromatography (RP-HPLC), and molecular modeling (PM6). Spectrophotometric measurements in aqueous solutions were performed at different temperatures. The stoichiometry of the complex is 1:1 and its apparent formation constant (Kc) is 3846 M−1 at 30 °C. Temperature dependence of Kc values revealed that both enthalpy (ΔH° = −10.58 kJ/mol) and entropy changes (ΔS° = 33.76 J/K mol) are favorable for the inclusion process in an aqueous medium. Encapsulation was also investigated using RP-HPLC (C18 column) with different mobile-phase compositions, to which β-CD was added. The apparent formation constants in MeOH–H2O (KF) were dependent of the proportion of the mobile phase employed (50:50, 55:45, 60:40 and 65:35, v/v). The KF values were 419 M−1 (50% MeOH) and 166 M−1 (65% MeOH) at 30 °C. The thermodynamic parameters of the complex in an aqueous MeOH medium indicated that this process is largely driven by enthalpy change (ΔH° = −27.25 kJ/mol and ΔS° = −45.12 J/K mol). The results of the study carried out with the PM6 semiempirical method showed that the energetically most favorable structure for the formation of the complex is the ‘head up’ orientation.  相似文献   

8.
Prislan I  Khutsishvili I  Marky LA 《Biochimie》2011,93(8):1341-1350
In the presence of specific metal ions, DNA oligonucleotides containing guanine repeat sequences can adopt G-quadruplex structures. In this work, we used a combination of spectroscopic and calorimetric techniques to investigate the conformation and unfolding thermodynamics of the K+-form of five G-quadruplexes with sequences: d(G2T2G2TGTG2T2G2), G2, d(G3T2G3TGTG3T2G3), G3, their analogs where T is replaced with U, G2-U and G3-U, and r(G2U2G2UGUG2U2G2), rG2. These G-quadruplexes show CD spectra characteristic of the “chair” conformation (G2 and G2-U), or “basket” conformation (rG2); or a mixture of these two conformers (G3 and G3-U). Thermodynamic profiles show that the favorable folding of each G-quadruplex results from the typical compensation of a favorable enthalpy and unfavorable entropy contributions. G-quadruplex stability increase in the following order (in ΔG°20): rG2 (−1.3 kcal/mol) < G2 < G2-U <G3-U (chair) < G3 (chair) <G3-U (basket) < G3 (basket) (−8.6 kcal/mol), due to favorable enthalpy contribution from the stacking of G-quartets.We used ITC to determine thermodynamic binding profiles for the interaction of the minor groove ligands, netropsin and distamycin, with each G-quadruplex. Both ligands bind with high exothermic enthalpies (∼−10.8 kcal/mol), 1:1 stoichiometries, and weak affinities (∼8 × 104 M−1). The similarity of the binding thermodynamic profiles, together with the absence of induced Cotton effects, indicates a surface or outside binding mode. We speculate that the top and bottom surfaces of the G-quadruplex comprise the potential MGBL binding sites, where the ligand lies on the surface forming van der Waals interactions with the guanines of the G-quartets and loop nucleotides.  相似文献   

9.
A detailed investigation on the oxidation of aqueous sulfite and aqueous potassium hexacyanoferrate(II) by the title complex ion has been carried out using the stopped-flow technique over the ranges, 0.01≤[S(IV)]T≤0.05 mol dm−3, 4.47≤pH≤5.12, and 24.9≤θ≤37.6 °C and at ionic strength 1.0 mol dm−3 (NaNO3) for aqueous sulfite and 0.01≤[Fe(CN)6 4−]≤0.11 mol dm−3, 4.54≤pH≤5.63, and 25.0≤θ≤35.3 °C and at ionic strength 1.0 or 3.0 mol dm−3 (NaNO3) for the hexacyanoferrate(II) ion. Both redox processes are dependent on pH and reductant concentration in a complex manner, that is, for the reaction with aqueous sulfite, kobs={(k1K1K2K3+k2K1K4[H+])[S(IV)]T]/([H+]2+K1[H+]+K1K2) and for the hexacyanoferrate(II) ion, kobs={(k1K3K4K5+k2K3K6[H+])[Fe(CN)6 4−]T)/([H+]2+K3[H+]+K3K4). At 25.0 °C, the value of k1′ (the composite of k1K3) is 0.77±0.07 mol−1 dm3 s−1, while the value of k2′ (the composite of k2K4) is (3.78±0.17)×10−2 mol−1 dm3 s−1 for aqueous sulfite. For the hexacyanoferrate(II) ion, k1′ (the composite of k1K5) is 1.13±0.01 mol−1 dm3 s−1, while the value of k2′ (the composite of k2K6) is 2.36±0.05 mol−1 dm3 s−1 at 25.0 °C. In both cases there was reduction of the cobalt(III) centre to cobalt(II), but there was no reduction of the molybdenum(VI) centre. k22, the self-exchange rate constant, for aqueous sulfite (as SO3 2−) was calculated to be 5.37×10−12 mol−1 dm3 s−1, while for Fe(CN)6 4−, it was calculated to be 1.10×109 mol−1 dm3 s−1 from the Marcus equations.  相似文献   

10.
Electrostatic interactions have a central role in some biological processes, such as recognition of charged ligands by proteins. We characterized the binding energetics of yeast triosephosphate isomerase (TIM) with phosphorylated inhibitors 2-phosphoglycollate (2PG) and phosphoglycolohydroxamate (PGH). We determined the thermodynamic parameters of the binding process (Kb, ΔGb, ΔHb, ΔSb and ΔCp) with different concentrations of NaCl, using fluorimetric and calorimetric titrations in the conventional mode of ITC and a novel method, multithermal titration calorimetry (MTC), which enabled us to measure ΔCp in a single experiment. We ruled out specific interactions of Na+ and Cl- with the native enzyme and did not detect significant linked protonation effects upon the binding of inhibitors. Increasing ionic strength (I) caused Kb, ΔGb and ΔHb to become less favorable, while ΔSb became less unfavorable. From the variation of Kb with I, we determined the electrostatic contribution of TIM−2PG and TIM−PGH to ΔGb at I = 0.06 M and 25 °C to be 36% and 26%, respectively. The greater affinity of PGH for TIM is due to a more favorable ΔHb compared to 2PG (by 19-24 kJ mol-1 at 25 °C). This difference is compatible with PGH establishing up to five more hydrogen bonds with TIM. Both binding ΔCps were negative, and less negative with increasing ionic strength. ΔCps at I = 0.06 M were much more negative than predicted by surface area models. Water molecules trapped in the interface when ligands bind to protein could explain the highly negative ΔCps. Thermodynamic binding functions for TIM−2PG changed more with ionic strength than those for TIM−PGH. This greater dependence is consistent with linked, but compensated, protonation equilibriums yielding the dianionic species of 2PG that binds to TIM, process that is not required for PGH.  相似文献   

11.
Cartilage Acidic Protein 2 (CRTAC2) is a novel protein present from prokaryotes to vertebrates with abundant expression in the teleost fish pituitary gland and an isoform of CRTAC1, a chondrocyte marker in humans. The two proteins are non-integrins containing N-terminal integrin-like Ca2 +-binding motifs and their structure and function remain to be assigned. Structural studies of recombinant sea bream (sb)CRTAC2 revealed it is composed of 8.8% α-helix, 33.4% β-sheet and 57.8% unordered protein. sbCRTAC2 bound Ca2 + with high affinity (Kd = 1.46 nM) and favourable Gibbs free energy (?G = − 12.4 kcal/mol). The stoichiometry for Ca2 + bound to sbCRTAC2 at saturation indicated six Ca2 + ligand-binding sites exist per protein molecule. No conformational change in sbCRTAC2 occurred in the presence of Ca2 +. Fluorescence emission revealed that the tertiary structure of the protein is hyperthermostable between 25 °C and 95 °C and the fully unfolded state is only induced by chemical denaturing (4 M GndCl). sbCRTAC has a widespread tissue distribution and is present as high molecular weight aggregates, although strong reducing conditions promote formation of the monomer. sbCRTAC2 promotes epithelial cell outgrowth in vitro suggesting it may share functional homology with mammalian CRTAC1, recently implicated in cell–cell and cell–matrix interactions.  相似文献   

12.
Microbes and plants synthesize essential branched-chain amino acids (BCAAs) such as valine, leucine, and isoleucine via a common biosynthetic pathway in which the first reaction is catalyzed by acetohydroxyacid synthase (AHAS, EC 4.1.3.18). Recently, AHAS was identified as a potential anti bacterial target. To help find an effective inhibitor that could act as an antibacterial compound, we cloned and characterized the catalytic subunit (CSU) of Pseudomonas aeruginosa AHAS, and found four potent inhibitors through chemical library screening. The ilvI gene of P. aeruginosa encodes a 65-kDa AHAS protein, consistent with the size of the purified enzyme on SDS-PAGE. Enzyme kinetics showed that the enzyme has a Km of 14.2 mM and a specific activity of 0.12 U/mg. Enzyme activity was optimum at a temperature of 37 °C and a pH of 7.5. The Kd for thiamine diphosphate (ThDP) was 89.92 ± 17.9 μM, as determined by fluorescence quenching. The cofactor activation constants (Ks) for ThDP and (Kc) for Mg2+ were 0.6 ± 0.1 and 560.8 ± 7.4 μM, respectively. Further, we determined that AVS2087, AVS2093, AVS2236, and AVS2387 compounds are potent inhibitors of the catalytic subunit of P. aeruginosa AHAS. These compounds inhibit nearly 100% of AHAS activity, with IC50 values of 1.19 μM, 5.0 nM, 25 nM, and 13 nM, respectively. Compound AVS2093 showed growth inhibition with a minimal inhibitory concentration (MIC) of 742.9 μg/ml against P. aeruginosa strain ATCC 9027. Furthermore, these findings were supported by molecular docking studies with the AVS compounds against P. aeruginosa AHAS in which AVS2093 showed minimum binding energy (−7.8 kJ/mol) by interacting with the receptor through a single hydrogen bond of 2.873 Å. Correlation of AVS2093 activity with P. aeruginosa AHAS cell growth inhibition suggested that AHAS might serve as a target protein for the development of novel antibacterial therapeutics. Results of the current study provide an impetus to further evaluate the potency of these inhibitors against pathogenic P. aeruginosa strains in vivo and to design more potent antibacterial agents based on these AVS inhibitors.  相似文献   

13.
Lipopolysaccharide (LPS), the major constituent of the outer membrane of Gram-negative bacteria, is the very first site of interactions with the antimicrobial peptides. In this work, we have determined a solution conformation of melittin, a well-known membrane active amphiphilic peptide from honey bee venom, by transferred nuclear Overhauser effect (Tr-NOE) spectroscopy in its bound state with lipopolysaccharide. The LPS bound conformation of melittin is characterized by a helical structure restricted only to the C-terminus region (residues A15-R24) of the molecule. Saturation transfer difference (STD) NMR studies reveal that several C-terminal residues of melittin including Trp19 are in close proximity with LPS. Isothermal titration calorimetry (ITC) data demonstrates that melittin binding to LPS or lipid A is an endothermic process. The interaction between melittin and lipid A is further characterized by an equilibrium association constant (Ka) of 2.85 × 106 M− 1 and a stoichiometry of 0.80, melittin/lipid A. The estimated free energy of binding (ΔG0), − 8.8 kcal mol− 1, obtained from ITC experiments correlates well with a partial helical structure of melittin in complex with LPS. Moreover, a synthetic peptide fragment, residues L13-Q26 or mel-C, derived from the C-terminus of melittin has been found to contain comparable outer membrane permeabilizing activity against Escherichia coli cells. Intrinsic tryptophan fluorescence experiments of melittin and mel-C demonstrate very similar emission maxima and quenching in presence of LPS micelles. The Red Edge Excitation Shift (REES) studies of tryptophan residue indicate that both peptides are located in very similar environment in complex with LPS. Collectively, these results suggest that a helical conformation of melittin, at its C-terminus, could be an important element in recognition of LPS in the outer membrane.  相似文献   

14.
Our goal was to gain a better understanding of the contribution of hydrophobic interactions to protein stability. We measured the change in conformational stability, Δ(ΔG), for hydrophobic mutants of four proteins: villin headpiece subdomain (VHP) with 36 residues, a surface protein from Borrelia burgdorferi (VlsE) with 341 residues, and two proteins previously studied in our laboratory, ribonucleases Sa and T1. We compared our results with those of previous studies and reached the following conclusions: (1) Hydrophobic interactions contribute less to the stability of a small protein, VHP (0.6 ± 0.3 kcal/mol per -CH2- group), than to the stability of a large protein, VlsE (1.6 ± 0.3 kcal/mol per -CH2- group). (2) Hydrophobic interactions make the major contribution to the stability of VHP (40 kcal/mol) and the major contributors are (in kilocalories per mole) Phe18 (3.9), Met13 (3.1), Phe7 (2.9), Phe11 (2.7), and Leu21 (2.7). (3) Based on the Δ(ΔG) values for 148 hydrophobic mutants in 13 proteins, burying a -CH2- group on folding contributes, on average, 1.1 ± 0.5 kcal/mol to protein stability. (4) The experimental Δ(ΔG) values for aliphatic side chains (Ala, Val, Ile, and Leu) are in good agreement with their ΔGtr values from water to cyclohexane. (5) For 22 proteins with 36 to 534 residues, hydrophobic interactions contribute 60 ± 4% and hydrogen bonds contribute 40 ± 4% to protein stability. (6) Conformational entropy contributes about 2.4 kcal/mol per residue to protein instability. The globular conformation of proteins is stabilized predominantly by hydrophobic interactions.  相似文献   

15.
The temperature dependence for the hydrolysis of both 4-methylumbelliferyl-α-l-fucoside and p-nitrophenyl-α-l-fucoside was determined for purified α-l-fucosidase (EC 3.2.1.51) from human placenta. The inhibition of the enzymatic reaction by l-fucose was also studied using the first of these two substrates at different temperatures. The thermodynamic parameters calculated from the pKm were for the 4-methylumbelliferyl-conjugate ΔF = ?6.6 kcal/mol, ΔH = ?8.5 kcal/mol, and ΔS = ?6.3 e.u. and for the p-nitrophenylconjugate ΔF = ?5.6 kcal/mol, ΔH = ?12.2 kcal/mol, and ΔS = ?21.1 e.u. The thermodynamic parameters for l-fucose were ΔH = ?12.4 kcal/mol and ΔS = ?20.1 e.u. The lower exothermicity and negative entropy calculated for the 4-methylumbelliferyl substrate compared to the thermodynamic parameters calculated for the p-nitrophenyl substrate and l-fucose suggest the existence of a secondary hydrophobic binding site for the 4-methylumbelliferyl moiety on the enzyme. The difference in the enthalpy for both substrates is also reflected in a difference in activation energy, being 15.8 kcal/mol for the 4-methylumbelliferyl substrate and 20.7 kcal/mol for the p-nitrophenyl substrate. From these results it may be concluded that altered kinetic properties of the enzyme could be the result of the binding of the “aglycone” moiety of the fluorogenic substrate to the enzyme.  相似文献   

16.
Exposure of winter rye leaves grown at 20°C and an irradiance of either 50 or 250 μmol m−2 s−1 to high light stress (1600 μmol m−2 s−1, 4 h) at 5°C resulted in photoinhibition of PSI measured in vivo as a 34% and 31% decrease in ΔA820/A820 (P700+). The same effect was registered in plants grown at 5°C and 50 μmol m−2 s−1. This was accompanied by a parallel degradation of the PsaA/PsaB heterodimer, increase of the intersystem e pool size as well as inhibition of PSII photochemistry measured as Fv/Fm. Surprisingly, plants acclimated to high light (800 μmol m−2 s−1) or to 5°C and moderate light (250 μmol m−2 s−1) were fully resistant to photoinhibition of PSI and did not exhibit any measurable changes at the level of PSI heterodimer abundance and intersystem e pool size, although PSII photochemistry was reduced to 66% and 64% respectively. Thus, we show for the first time that PSI, unlike PSII, becomes completely resistant to photoinhibition when plants are acclimated to either 20°C/800 μmol m−2 s−1 or 5°C/250 μmol m−2 s−1 as a response to growth at elevated excitation pressure. The role of temperature/light dependent acclimation in the induction of selective tolerance to PSI photoinactivation is discussed.  相似文献   

17.
A kinetic study of the rate of pyruvate reduction by goldfish LDH-M4 (the homotetrameric form of lactate dehydrogenase which predominates in skeletal muscle) provided an analysis of the effects of pH and temperature on v (reaction velocity) and estimates of how temperature might affect catalysis in vivo, where the physiological pH regulation imposes a temperature coefficient of ?0.015 to ?0.020 pH unit/ °C. Consistent with published data for other LDHs, (i) V (maximum reaction velocity) was pH insensitive over a physiological pH range, (ii) the Km for pyruvate, KP, was sensitive to both pH and temperature, and (iii) the Km for NADH and the dissociation constant for NADH were both sensitive to temperature, but not to pH. V approximately doubled with each 10 °C (Ea = 11.7 kcal/mol). The effects of pH and temperature on KP were consistent with two enthalpic contributions, an ionization enthalpy (ΔHi°) of 7.2 kcal/mol (probably a histidine imidazole), and an enthalpy (ΔHSO) of 5.8 kcal/mol for the combination of pyruvate with the nonionized (pH ? pK′) LDH-NADH complex. When the pH was varied according to the physiological temperature coefficient, v was more sensitive to temperature than for conditions of constant pH, the usual design of kinetic experiments. This effect was due to the decreased temperature sensitivity of KP caused by partial concellation of the ΔHi° effect by the pH regulation: dpHdT ? dpK′dT. At constant pH, on the other hand, KP increased strongly with temperature and had the effect of offsetting (especially at higher pH values) the large increases in V. It was suggested that the magnitudes of ΔHi° and ΔHSO might have been important in the evolution of LDHs and other enzymes of cold-blooded animals.  相似文献   

18.
Sickle hemoglobin forms long, multistranded polymers that account for the pathophysiology of the disease. The molecules in these polymers make significant contacts along the polymer axis (i.e., axial contacts) as well as making diagonally directed contacts (i.e., lateral contacts). The axial contacts do not engage the mutant β6 Val and its nonmutant receptor region on an adjacent molecule, in contrast to the lateral contacts which do involve the mutation site. We have studied the association process by elastic light scattering measurements as a function of temperature, concentration, and primary and quaternary structure, employing an instrument of our own construction. Even well below the solubility for polymer formation, we find a difference between the association behavior of deoxy sickle hemoglobin molecules (HbS), which can polymerize at higher concentration, in comparison to COHbS, COHbA, or deoxygenated Hemoglobin A (HbA), none of which have the capacity to form polymers. The nonpolymerizable species are all quite similar to one another, and show much less association than deoxy HbS. We conclude that axial contacts are significantly weaker than the lateral ones. All the associations are entropically favored, and enthalpically disfavored, typical of hydrophobic interactions. For nonpolymerizable Hemoglobin, ΔHo was 35 ± 4 kcal/mol, and ΔS was 102.7 ± 0.5 cal/(mol−K). For deoxyHbS, ΔHo was 19 ± 2 kcal/mol, and ΔS was 56.9 ± 0.5 cal/(mol−K). The results are quantitatively consistent with the thermodynamics of polymer assembly, suggesting that the dimer contacts and polymer contacts are very similar, and they explain a previously documented significant anisotropy between bending and torsional moduli. Unexpectedly, the results also imply that a substantial fraction of the hemoglobin has associated into dimeric species at physiological conditions.  相似文献   

19.
Ferric human serum heme-albumin (heme-HSA) shows a peculiar nuclear magnetic relaxation dispersion (NMRD) behavior that allows to investigate structural and functional properties. Here, we report a thermodynamic analysis of NMRD profiles of heme-HSA between 20 and 60 °C to characterize its hydration. NMRD profiles, all showing two Lorentzian dispersions at 0.3 and 60 MHz, were analyzed in terms of modulation of the zero field splitting tensor for the S = 5/2 manifold. Values of correlation times for tensor fluctuation (τv) and chemical exchange of water molecules (τM) show the expected temperature dependence, with activation enthalpies of −1.94 and −2.46 ± 0.2 kJ mol−1, respectively. The cluster of water molecules located in the close proximity of the heme is progressively reduced in size by increasing the temperature, with Δ= 68 ± 28 kJ mol−1 and Δ= 200 ± 80 J mol−1 K−1. These results highlight the role of the water solvent in heme-HSA structure-function relationships.  相似文献   

20.
Oxyopinins (Oxki1 and Oxki2) are antimicrobial peptides isolated from the crude venom of the wolf spider Oxyopes kitabensis. The effect of oxyopinins on lipid bilayers was investigated using high-sensitivity titration calorimetry and 31P solid-state NMR spectroscopy. High-sensitivity titration calorimetry experiments showed that the binding of oxyopinins was exothermic, and the binding enthalpies (ΔH) to 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylcholine (POPC) small unilamellar vesicles (SUVs) were − 18.1 kcal/mol and − 15.0 kcal/mol for Oxki1 and Oxki2, respectively, and peptide partition coefficient (Kp) was found to be 3.9 × 103 M− 1. 31P NMR spectra of 1,2-dielaidoyl-sn-glycero-3-phosphoethanolamine (DEPE) membranes in the presence of oxyopinins indicated that they induced a positive curvature in lipid bilayers. The induced positive curvature was stronger in the presence of Oxki2 than in the presence of Oxki1. 31P NMR spectra of phosphaditylcholine (PC) membranes in the presence of Oxki2 showed that Oxki2 produced micellization of membranes at low peptide concentrations, but unsaturated PC membranes or acidic phospholipids prevented micellization from occurring. Furthermore, 31P NMR spectra using membrane lipids from E. coli suggested that Oxki1 was more disruptive to bacterial membranes than Oxki2. These results strongly correlate to the known biological activity of the oxyopinins.  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