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 500 毫秒
1.
Leesch VW  Bujons J  Mauk AG  Hoffman BM 《Biochemistry》2000,39(33):10132-10139
Cytochrome c peroxidase (CcP) can bind as many as two cytochrome c (Cc) molecules in an electrostatic complex. The location of the two binding domains on CcP has been probed by photoinduced interprotein electron transfer (ET) between zinc-substituted horse cytochrome c (ZnCc) and CcP with surface charge-reversal mutations and by isothermal titration calorimetry (ITC). These results, which are the first experimental evidence for the location of domain 2, indicate that the weak-binding domain includes residues 146-150 on CcP. CcP(E290K) has a charge-reversal mutation in the tight-binding domain, which should weaken binding, and it weakens the 1:1 complex; K(1) decreases 20-fold at 18 mM ionic strength. We have employed two mutations to probe the proposed location for the weak-binding domain on the CcP surface: (i) D148K, a "detrimental" mutation with a net (+2) change in the charge of CcP, and (ii) K149E, a "beneficial" mutation with a net (-2) change in the charge. The interactions between FeCc and CcP(WT and K149E) also have been studied with ITC. The CcP(D148K) mutation causes no substantial change in the 2:1 binding but an increase in the reactivity of the 2:1 complex. The latter can be interpreted as a long-range influence on the heme environment or, more likely, the enhancement of a minority subset of binding conformations with favorable pathways for ET. CcP(K149E) has a charge-reversal mutation in the weak-binding domain that produces a substantial increase in the 2:1 binding constant as measured by both quenching and ITC. For the 1:1 complex of CcP(WT), DeltaG(1) = -8.2 kcal/mol (K(1) = 1.3 x 10(6) M(-)(1)), DeltaH(1) = +2.7 kcal/mol, and DeltaS(1) = +37 cal/K.mol at 293 K; for the second binding stage, K(2) < 5 x 10(3) M(-)(1), but accurate thermodynamic parameters were not obtained. For the 1:1 complex of CcP(K149E), DeltaG(1) = -8.5 kcal/mol (K(1) = 2 x 10(6) M(-)(1)), DeltaH(1) = +2. 0 kcal/mol, and DeltaS(1) = +36 cal/K.mol; for the second stage, DeltaG(2) = -5.5 kcal/mol (K(1) = 1.3 x 10(4) M(-)(1)), DeltaH(2) = +2.9 kcal/mol, and DeltaS(2) = +29 cal/K.mol.  相似文献   

2.
Kleeb AC  Kast P  Hilvert D 《Biochemistry》2006,45(47):14101-14110
Prephenate dehydratase (PDT) is an important but poorly characterized enzyme that is involved in the production of L-phenylalanine. Multiple-sequence alignments and a phylogenetic tree suggest that the PDT family has a common structural fold. On the basis of its sequence, the PDT from the extreme thermophile Methanocaldococcus jannaschii (MjPDT) was chosen as a promising representative of this family for pursuing structural and functional studies. The corresponding pheA gene was cloned and expressed in Escherichia coli. It encodes a monofunctional and thermostable enzyme with an N-terminal catalytic domain and a C-terminal regulatory ACT domain. Biophysical characterization suggests a dimeric (62 kDa) protein with mixed alpha/beta secondary structure elements. MjPDT unfolds in a two-state manner (Tm = 94 degrees C), and its free energy of unfolding [DeltaGU(H2O)] is 32.0 kcal/mol. The purified enzyme catalyzes the conversion of prephenate to phenylpyruvate according to Michaelis-Menten kinetics (kcat = 12.3 s-1 and Km = 22 microM at 30 degrees C), and its activity is pH-independent over the range of pH 5-10. It is feedback-inhibited by L-phenylalanine (Ki = 0.5 microM), but not by L-tyrosine or L-tryptophan. Comparison of its activation parameters (DeltaH(++)= 15 kcal/mol and DeltaS(++)= -3 cal mol-1 K-1) with those for the spontaneous reaction (DeltaH(++) = 17 kcal/mol and DeltaS(++)= -28 cal mol-1 K-1) suggests that MjPDT functions largely as an entropy trap. By providing a highly preorganized microenvironment for the dehydration-decarboxylation sequence, the enzyme may avoid the extensive solvent reorganization that accompanies formation of the carbocationic intermediate in the uncatalyzed reaction.  相似文献   

3.
The thermal inactivation of immobilized cholinesterase enzymes (ChE) in solid matrices where the protein unfolding is blocked was studied, thus enabling investigation of the kinetics of the inactivation process directly from the native structure to the inactivated state. The thermal inactivation of butyrylcholinesterase (BChE), recombinant human acetylcholinesterase (rHuAChE), and eel acetylcholinesterase (AChE) enzymes was studied in dry films composed of poly(vinyl pyrollidone) (PVP), bovine serum albumin (BSA) and trehalose at 60 degrees -120 degrees C. The kinetics follows a bi-exponential decay equation representing a combination of fast and slow processes. The activation enthalpy DeltaH(#) and the activation entropy DeltaS(#) for each of the three enzymes have been evaluated. The values of DeltaH(#) for the fast process and for the slow process of BChE are 33+/-3, and 28+/-2 kcal/mol, respectively, and the values of DeltaS(#) are 0.84+/-0.04, and -18.2+/-0.5 cal/deg, respectively. The appropriate value of DeltaH(#) for rHuAChE is 26+/-2 Kcal/mol, for both processes and the values of DeltaS(#) are -17.6+/-0.9, and -23.0+/-0.9 cal/deg, respectively. Similarly, the values of DeltaH(#) for eelAChE are 30+/-3, 31+/-1 kcal/mol, and the values of DeltaS(#) are -6.7+/-0.5, -9.1+/-0.2 cal/deg respectively.  相似文献   

4.
The formal equilibrium reduction potentials of recombinant electron transport protein, rubredoxin (MW = 7500 Da), from both the mesophilic Clostridium pasteurianum (Topt = 37 degrees C) and hyperthermophilic Pyrococcus furiosus (Topt = 95 degrees C) were recorded as a function of pressure and temperature. Measurements were made utilizing a specially designed stainless steel electrochemical cell that easily maintains pressures between 1 and 600 atm and a temperature-controlled cell that maintains temperatures between 4 and 100 degrees C. The reduction potential of P. furiosus rubredoxin was determined to be 31 mV at 25 degrees C and 1 atm, -93 mV at 95 degrees C and 1 atm, and 44 mV at 25 degrees C and 400 atm. Thus, the reduction potential of P. furiosus rubredoxin obtained under standard conditions is likely to be dramatically different from the reduction potential obtained under its normal operating conditions. Thermodynamic parameters associated with electron transfer were determined for both rubredoxins (for C. pasteurianum, DeltaV degrees = -27 mL/mol, DeltaS degrees = -36 cal K-1 mol-1, and DeltaH degrees = -10 kcal/mol, and for P. furiosus, DeltaV degrees = -31 mL/mol, DeltaS degrees = -41 cal K-1 mol-1, and DeltaH degrees = -13 kcal/mol) from its pressure- and temperature-reduction potential profiles. The thermodynamic parameters for electron transfer (DeltaV degrees, DeltaS degrees, and DeltaH degrees ) for both proteins were very similar, which is not surprising considering their structural similarities and sequence homology. Despite the fact that these two proteins exhibit dramatic differences in thermostability, it appears that structural changes that confer dramatic differences in thermostability do not significantly alter electron transfer reactivity. The experimental changes in reduction potential as a function of pressure and temperature were simulated using a continuum dielectric electrostatic model (DELPHI). A reasonable estimate of the protein dielectric constant (epsilonprotein) of 6 for both rubredoxins was determined from these simulations. A discussion is presented regarding the analysis of electrostatic interaction energies of biomolecules through pressure- and temperature-controlled electrochemical studies.  相似文献   

5.
Photoacoustic calorimetry has been utilized to probe the thermodynamics accompanying photodissociation of the CO mixed valence form of bovine heart cytochrome c oxidase (COMV CcO). At pH's below 9 photolysis of the COMV CcO results in three kinetic phases with the first phase occurring faster than the time resolution of the instrument (i.e., < approximately 50 ns), a second phase occurring with a lifetime of approximately 100 ns and a third phase occurring with a lifetime of approximately 2 micros. The corresponding volume and enthalpy changes for these processes are: DeltaH1, DeltaV1 = +79 +/- 10 kcal mol(-1), +9 +/- 1 mL mol(-1); DeltaH2, DeltaV2 = -79 +/- 5 kcal mol(-1), -9 +/- 2 mL mol(-1); DeltaH3, DeltaV3 = +54 +/- 7 kcal mol(-1), +8 +/- 1 mL mol(-1). At pH's above 9 only one phase is observed, a prompt phase occurring in < 50 ns. The overall volume change is negligible above pH 9 and the enthalpy change is +29 +/- 5 kcal mol(-1). The data are consistent with the prompt phase being associated with CO-Fe(a3) bond cleavage, CO-CuB+ bond formation, Fe(a3) low-spin to high-spin transition and fast electron transfer (ET) from heme a3 to heme a followed by proton transfer from Glu242 to Arg38 on an approximately 100 ns timescale. The slow phase is likely a combination of CO thermal dissociation from CuB and additional ET between heme a3 to heme a. Interestingly, this phase is not evident above pH 9 suggesting linkage between CO dissociation/ET and the protonation state of a group or groups near the binuclear center.  相似文献   

6.
7.
The self-association of Escherichia coli alpha-ketoglutarate dehydrogenase complex (KGDC) purified by a column Chromatographic technique, was characterized by light-scattering photometry. The complex adopts a solution conformation somewhat larger than that observed in the electron microscope. The evidence suggests a nonideal indefinite self-association model for KGDC in KCl, phosphate buffer. The KGDC monomer has a molecular charge of about -3 x 10(2) at neutral pH. The self-association is promoted by increasing KCl concentrations, pH (in the range from 6.3 to 7.4) and temperature (from 20 to 30 degrees C). The effects of pH changes suggest a release of protons during the self-association and a minor 'preferential' interaction of phosphate ions. For the association of one monomer to the aggregate at neutral pH and 25 degrees C. DeltaG degrees = -7.8 kcal mol(-1). DeltaH degrees = 24 kcal mol(-1) and DeltaS degrees = 1.1 x 10(2) cal mol(-1) K(-1). These data indicate that hydrophobic interactions drive the association. Thermodynamically, the self-association of KGDC is a complex phenomenon and may serve to stabilize the enzyme complex in solution.  相似文献   

8.
Horse heart ferric cytochrome c was investigated by the following three methods: (I) Light absorption spectrophotometry at 23 degrees C and 77 degrees K; (II) Electron paramagnetic resonance (EPR) spectroscopy at 20 degrees K; (III) Precise equilibrium measurements of ferric cytochrome c with azide and imidazole between 14.43 and 30.90 degrees C. I and II have demonstrated that: (1) Ferric cytochrome c azide and imidazole complexes were in the purely low spin state between 20 degrees K and 23 degrees C; (2) The energy for the three t2g orbitals calculated in one hole formalism shows that azide or imidazole bind to the heme iron in a similar manner to met-hemoglobin azide or imidazole complexes, respectively. III has demonstrated that: (1) The change of standard enthalpy and that of standard entropy were -2.3 kcal/mol and -1.6 cal/mol per degree for the azide complex formation, and -1.4 kcal/mol and 2.9 cal/mol per degree for the imidazole complex formation. (2) A linear relationship between the change of entropy and that of enthalpy was observed for the above data for the cyanide complex formation. The complex formation of ferric cytochrome c was discussed based on the results of X-ray crystallographic studies compared with hemoglobin and myoglobin.  相似文献   

9.
NMR study of the alkaline isomerization of ferricytochrome c   总被引:1,自引:0,他引:1  
X L Hong  D W Dixon 《FEBS letters》1989,246(1-2):105-108
The pH-induced isomerization of horse heart cytochrome c has been studied by 1H NMR. We find that the transition occurring in D2O with a pKa measured as 9.5 +/- 0.1 is from the native species to a mixture of two basic forms which have very similar NMR spectra. The heme methyl peaks of these two forms have been assigned by 2D exchange NMR. The forward rate constant (native to alkaline cytochrome c) has a value of 4.0 +/- 0.6 s-1 at 27 degrees C and is independent of pH; the reverse rate constant is pH-dependent. The activation parameters are delta H not equal to = 12.8 +/- 0.8 kcal.mol1, delta S not equal to = -12.9 +/- 2.0 e.u. for the forward reaction and delta H not equal to = 6.0 +/- 0.3 kcal.mol-1, delta S not equal to = -35.1 +/- 1.3 e.u. for the reverse reaction (pH* = 9.28). delta H degree and delta S degree for the isomerization are 6.7 +/- 0.6 kcal.mol-1 and 21.9 +/- 1.0 e.u., respectively.  相似文献   

10.
The peripheral subunit-binding domain (PSBD) of the dihydrolipoyl acetyltransferase (E2, EC 2.3.1.12) binds tightly but mutually exclusively to dihydrolipoyl dehydrogenase (E3, EC 1.8.1.4) and pyruvate decarboxylase (E1, EC 1.2.4.1) in the pyruvate dehydrogenase multienzyme complex of Bacillus stearothermophilus. Isothermal titration calorimetry (ITC) experiments demonstrated that the enthalpies of binding (DeltaH degrees ) of both E3 and E1 with the PSBD varied with salt concentration, temperature, pH, and buffer composition. There is little significant difference in the free energies of binding (DeltaG degrees = -12.6 kcal/mol for E3 and = -12.9 kcal/mol for E1 at pH 7.4 and 25 degrees C). However, the association with E3 was characterized by a small, unfavorable enthalpy change (DeltaH degrees = +2.2 kcal/mol) and a large, positive entropy change (TDeltaS degrees = +14.8 kcal/mol), whereas that with E1 was accompanied by a favorable enthalpy change (DeltaH degrees = -8.4 kcal/mol) and a less positive entropy change (TDeltaS degrees = +4.5 kcal/mol). Values of DeltaC(p) of -316 cal/molK and -470 cal/molK were obtained for the binding of E3 and E1, respectively. The value for E3 was not compatible with the DeltaC(p) calculated from the nonpolar surface area buried in the crystal structure of the E3-PSBD complex. In this instance, a large negative DeltaC(p) is not indicative of a classical hydrophobic interaction. In differential scanning calorimetry experiments, the midpoint melting temperature (T(m)) of E3 increased from 91 degrees C to 97.1 degrees C when it was bound to PSBD, and that of E1 increased from 65.2 degrees C to 70.0 degrees C. These high T(m) values eliminate unfolding as a major source of the anomalous DeltaC(p) effects at the temperatures (10-37 degrees C) used for the ITC experiments.  相似文献   

11.
SecA is the dissociatable nucleotide and preprotein binding subunit of the bacterial translocase. The thermodynamics of nucleotide binding to soluble SecA at nucleotide binding site I were determined by isothermal titration calorimetry. Binding of ADP and non-hydrolyzable ATPgammaS is enthalpy-driven (DeltaH(0) of -14.44 and -5.56 kcal/mol, respectively), but is accompanied by opposite entropic contributions (DeltaS(0) of -18.25 and 9.55 cal/mol/K, respectively). ADP binding results in a large change in the heat capacity of SecA (DeltaC(p)=-780 cal/mol/K). It is suggested that ADP binding promotes the interaction between the two thermodynamically discernible domains of SecA which is accompanied by a shielding of hydrophobic surface from solvent.  相似文献   

12.
13.
Kinetic studies of the reduction of Pseudomonas aeruginosa ferricytochrome c551 by Fe(EDTA)2- have been made. The reaction was found to follow a second-order rate law: k 4.2 x 10(3) M(-1) s(-1) [25 degrees, micro0.1 M, pH 7.0 (phosphate)]; deltaH+/+ 3.2 kcal/ mol; AS+/+ -30 cal/mol-deg. The electrostatics-corrected self-exchange rate constant (k11 corr) calculated for cytochrome c551 based on the Fe(EDTA)2- cross reaction is 2 M(-1) s(-1), as compared to a value of 6 M(-1) s(-1) for horse heart cytochrome c. The close correspondence of the two k11 corr values is taken as an indication that the two proteins employ very similar electron transfer mechanisms in their reactions with Fe(EDTA)(2-). It is proposed that this mechanism involves reagent contact, but little protein conformational change, at the partially exposed heme edge.  相似文献   

14.
M Yang  D Liu  D W Bolen 《Biochemistry》1999,38(34):11216-11222
This work determines the ratio of DeltaH(vH) /DeltaH(cal) for staphylococcal nuclease (SN) denaturation in guanidine hydrochloride (GdnHCl) to test whether GdnHCl-induced denaturation is two-state. Heats of mixing of SN as a function of [GdnHCl] were determined at pH 7.0 and 25 degrees C. The resulting plot of DeltaH(mix) vs [GdnHCl] exhibits a sigmoid shaped curve with linear pre- and post-denaturational base lines. Extending the pre- and post-denaturational lines to zero [GdnHCl] gives a calorimetric DeltaH (DeltaH(cal)) of 24.1 +/- 1.0 kcal/mol, for SN denaturation in the limit of zero GdnHCl concentration. Guanidine hydrochloride-induced denaturation Gibbs energy changes in the limit of zero denaturant concentration (DeltaG degrees (N)(-)(D)) at pH 7. 0 were determined for SN from fluorescence measurements at fixed temperatures over the range from 15 to 35 degrees C. Analysis of the resulting temperature-dependent DeltaG degrees (N)(-)(D) data defines a van't Hoff denaturation enthalpy change (DeltaH(vH)) of 26. 4 +/- 2.8 kcal/mol. The model-dependent van't Hoff DeltaH(vH) divided by the model-independent DeltaH(cal) gives a ratio of 1.1 +/- 0.1 for DeltaH(vH)/DeltaH(cal), a result that rules out the presence of thermodynamically important intermediate states in the GdnHCl-induced denaturation of SN. The likelihood that GdnHCl-induced SN denaturation involves a special type of two-state denaturation, known as a variable two-state process, is discussed in terms of the thermodynamic implications of the process.  相似文献   

15.
Barceló F  Portugal J 《FEBS letters》2004,576(1-2):68-72
The antitumor drug elsamicin A contains a coumarin-related chartarin chromophore that intercalates into DNA. It differs from other related molecules in its disaccharide moiety, which bears an amino sugar. Its binding to DNA was analyzed using isothermal titration calorimetry and UV thermal denaturation, and characterized thermodynamically. For the association of elsamicin A with DNA we found DeltaG degrees = -8.6 kcal mol(-1), DeltaH = -10.4 kcal mol(-1), DeltaS = -6.1 cal mol(-1) K(-1), and Kobs = 2.8(+/- 0.2) x 10(6) M(-1) at 20 degrees C in 18 mM Na+. The contributions to the free energy of binding that lead to the DNA-elsamicin complex are compared with the binding to DNA of chartreusin, another chartarin-containing drug. The results are discussed in terms of the contributions of the disaccharide moieties into the strength of binding.  相似文献   

16.
Resonance Raman spectra of ferrous and ferric cytochrome c peroxidase and Compound ES and their pH dependences were investigated in resonance with Soret band. The Fe(IV) = O stretching Raman line of Compound ES was assigned to a broad band around 767 cm-1, which was shifted to 727 cm-1 upon 18O substitution. The 18O-isotopic frequency shift was recognized for Compound ES derived in H218O, but not in H216O. This clearly indicated occurrence of an oxygen exchange between the Fe(IV) = O heme and bulk water. The Fe(IV) = O stretching Raman band was definitely more intense and of higher frequency in D2O than in H2O as in Compound II of horseradish peroxidase, but in contrast with this its frequency was unaltered between pH 4 and 11. The Fe(II)-histidine stretching Raman line was assigned on the basis of the frequency shift observed for 54Fe isotopic substitution. From the intensity analysis of this band, the pKa of the heme-linked ionization of ferrocytochrome c peroxidase was determined to be 7.3. The Raman spectrum of ferricytochrome c peroxidase strongly suggested that the heme is placed under an equilibrium between the 5- and 6-coordinate high-spin structures. At neutral pH it is biased to the 5-coordinate structure, but at the acidic side of the transition of pKa = 5.5 the 6-coordinate heme becomes dominant. F- was bound to the heme iron at pH 6, but Cl- was bound only at acidic pH. Acidification by HNO3, H2SO4, CH3COOH, HBr, or HI resulted in somewhat different populations of the 5- and 6-coordinate forms when they were compared at pH 4.3. Accordingly, it is inferred that a water molecule which is suggested to occupy the sixth coordination position of the heme iron is not coordinated to the heme iron at pH 6 but that protonation of the pKa = 5.5 residue induces an appreciable structural change, allowing the coordination of the water molecule to the heme iron.  相似文献   

17.
Surfactin is a bacterial lipopeptide with powerful surfactant-like properties. High-sensitivity isothermal titration calorimetry was used to study the self association and membrane partitioning of surfactin. The critical micellar concentration (CMC), was 7.5 microM, the heat of micellization was endothermic with DeltaH(w-->m)(Su) = +4.0 kcal/mol, and the free energy of micellization DeltaG(O,w-->m)(Su) = -9.3 kcal/mol (25 degrees C; 100 mM NaCl; 10 mM TRIS, 1 mM EDTA; pH 8.5). The specific heat capacity of micellization was deduced from temperature dependence of DeltaH(w-->m)(Su) as DeltaC(w-->m)(P) = -250 +/- 10 cal/(mol.K). The data can be explained by combining the hydrophobicity of the fatty acyl chain with that of the hydrophobic amino acids. The membrane partition equilibrium was studied using small (30 nm) and large (100 nm) unilamellar POPC vesicles. At 25 degrees C, the partition coefficient, K, was (2.2 +/- 0.2) x 10(4) M(-1) for large vesicles leading to a free energy of DeltaG(O, w-->b)(Su) = -8.3 kcal/mol. The partition enthalpy was again endothermic, with DeltaH(w-->b)(Su) = 9 +/- 1 kcal/mol. The strong preference of surfactin for micelle formation over membrane insertion explains the high membrane-destabilizing activity of the peptide. For surfactin and a variety of non-ionic detergents, the surfactant-to-lipid ratio, inducing membrane solubilization, R(sat)(b), can be predicted by the simple relationship R(sat)(b) approximately K. CMC.  相似文献   

18.
19.
The oxidation of yeast cytochrome c peroxidase by hydrogen peroxide produces a unique enzyme intermediate, cytochrome c peroxidase Compound I, in which the ferric heme iron has been oxidized to an oxyferryl state, Fe(IV), and an amino acid residue has been oxidized to a radical state. The reduction of cytochrome c peroxidase Compound I by horse heart ferrocytochrome c is biphasic in the presence of excess ferrocytochrome c as cytochrome c peroxidase Compound I is reduced to the native enzyme via a second enzyme intermediate, cytochrome c peroxidase Compound II. In the first phase of the reaction, the oxyferryl heme iron in Compound I is reduced to the ferric state producing Compound II which retains the amino acid free radical. The pseudo-first order rate constant for reduction of Compound I to Compound II increases with increasing cytochrome c concentration in a hyperbolic fashion. The limiting value at infinite cytochrome c concentration, which is attributed to the intracomplex electron transfer rate from ferrocytochrome c to the heme site in Compound I, is 450 +/- 20 s-1 at pH 7.5 and 25 degrees C. Ferricytochrome c inhibits the reaction in a competitive manner. The reduction of the free radical in Compound II is complex. At low cytochrome c peroxidase concentrations, the reduction rate is 5 +/- 3 s-1, independent of the ferrocytochrome c concentration. At higher peroxidase concentrations, a term proportional to the square of the Compound II concentration is involved in the reduction of the free radical. Reduction of Compound II is not inhibited by ferricytochrome c. The rates and equilibrium constant for the interconversion of the free radical and oxyferryl forms of Compound II have also been determined.  相似文献   

20.
Effects of different end sequences on stability, circular dichroism spectra (CD), and enzyme binding properties were investigated for six 22-base pair, non-self-complementary duplex DNA oligomers. The center sequences of these deoxyoligonucleotides have 8-14 base pairs in common and are flanked on both sides by sequences differing in context and A-T content. Temperature-induced melting transitions monitored by differential scanning calorimetry (DSC) and ultraviolet absorbance were measured for the six duplexes in buffered 115 mM Na(+) solutions. Values of the melting transition enthalpy, DeltaH(cal), and entropy, DeltaS(cal), were obtained directly from DSC experiments. Melting transition parameters, DeltaH(vH) and DeltaS(vH), were also estimated from van't Hoff analysis of optical melting curves collected as a function of DNA concentration, assuming a two-state melting transition. Melting free energies (20 degrees C) of the six DNAs evaluated from DSC experiments ranged from -18.7 to -32.7 kcal/mol. van't Hoff estimates of the free energies ranged from -18.5 to -48.0 kcal/mol. With either method, the trends in free energy as a function of sequence were identical. Equilibrium binding by BamHI restriction endonuclease to the 22-base pair DNAs was also investigated. The central eight base pairs of all six molecules, 5'-A-GGATCC-A-3', contained a BamHI recognition sequence bounded by A-T base pairs. Magnesium free binding assays were performed by titering BamHI against a constant concentration of each of the deoxyoligonucleotide substrates and analyzing reaction products by gel retardation. Binding isotherms of the total amount of bound DNA versus protein concentration were constructed which provided semiquantitative estimates of the equilibrium dissociation constants for dissociation of BamHI from the six DNA oligomers. Dissociation constants ranged from 0.5 x 10(-)(9) to 12.0 x 10(-)(9) M with corresponding binding free energies of -12.5 to -10.6 (+/-0. 1) kcal/mol. An inverse relationship is found when binding and stability are compared.  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