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.
Dex-Benzedrine (known as d-Benzedrine or SAT) acts in dopamine receptors of central nerve cell system. In clinic, SAT is used to treat a variety of diseases; meanwhile, it has dependence and addiction. In order to investigate the pharmacology and addiction mechanisms of SAT as a medicine, in this paper, we have studied the structure of D3R complex protein with SAT, and based on which, using potential mean force with umbrella samplings and the simulated phospholipid bilayer membrane (or POPC bilayer membrane), the molecular dynamics simulation was performed to obtain free energy changes upon the trajectories for SAT moving along the molecular channels within D3R. The free energy change for SAT transmitting toward the outside of cell along the functional molecular channel within D3R is 83.5 kJ mol?1. The change of free energy for SAT to permeate into the POPC bilayer membrane along the protective molecular channel within D3R is 87.7 kJ mol?1. Our previous work gave that the free energy for Levo-Benzedrine (RAT) transmitting toward the outside of cell along the functional molecular channel within D3R is 91.4 kJ mol?1, while it is 117.7 kJ mol?1 for RAT to permeate into the POPC bilayer membrane along the protective molecular channel within D3R. The values of free energy suggest that SAT relatively prefers likely to pass through the functional molecular channel within D3R for increasing the release of dopamine molecules resulting in a variety of functional effects for SAT. The obtained results show that the pharmacology and addiction mechanisms of SAT as a drug are closely related to the molecular dynamics and mechanism for SAT transmitting along molecular channels within D3R.  相似文献   

2.
《Inorganica chimica acta》1988,154(2):209-214
The diastereoisomeric complex Δ-(+)-tris(cyclicO,O′, 1 (R), 2(R)(−)dimethylethylene dithiophosphato)chromium(III), was synthesized stereoselectively in tetrahydrofuran (THF) solution. The complex proves optically labile, [α]D=+106, in CHCl3, changing quickly to [α]D=+211. The CD spectra in THF enable us to characterize the complex and show a configuration inversion which gives the diastereoisomeric equilibrium Λ⇌Δ with an excess of the Λ-(R,R)(R,R)(R,R) diastereoisomeric form. The equilibrium constant K=0.86 at 25 °C is indicative of a different thermodynamic stability between the two diastereoisomers in THF solution, Λ-(R,R)> Δ-(R,R), δΔH°=1.5 kJ mol−1, δΔG°=0.3 kJ mol−1, δΔS°=4 J mol−1 K−1. The kinetic diastereoisomer Δ-(R,R)(R,R)(R,R) is stabilized in CHCl3, CH2Cl2, EtOH solvents where it is highly soluble and optically stable with a maximum negative chirality factor, g=−5×10−3, in CHCl3.  相似文献   

3.
T4 polynucleotide kinase (Pnk) is a bifunctional 5′-kinase/3′-phosphatase that aids in the repair of broken termini in RNA by converting 3′-PO4/5′-OH ends into 3′-OH/5′-PO4 ends, which are then sealed by RNA ligase. Here we have employed site-directed mutagenesis (introducing 31 mutations at 16 positions) to locate candidate catalytic residues within the 301 amino acid Pnk polypeptide. We found that alanine substitutions for Arg38 and Arg126 inactivated the 5′-kinase, but spared the 3′-phosphatase activity. Conservative substitutions of lysine or glutamine for Arg38 and Arg126 did not restore 5′-kinase activity. These results, together with previous mutational studies, highlight a constellation of five amino acids (Lys15, Ser16, Asp35, Arg38 and Arg126) that likely comprise the 5′-kinase active site. Four of these residues are conserved at the active sites of adenylate kinases (Adk), suggesting that Pnk and Adk are structurally and mechanistically related. We found that alanine substitutions for Asp165, Asp167, Arg176, Arg213, Asp254 and Asp278 inactivated the 3′-phosphatase, but spared the 5′-kinase. Conservative substitutions of asparagine or glutamate for Asp165, Asp167 and Asp254 did not revive the 3′-phosphatase activity, nor did lysine substitutions for Arg176 and Arg213. Glutamate in lieu of Asp278 partially restored activity, whereas asparagine had no salutary effect. Alanine substitutions for Arg246 and Arg279 partially inactivated the 3′-phosphatase; the conservative R246K change restored activity, whereas R279K had no benefit. The essential phosphatase residues Asp165 and Asp167 are located within a 165DxDxT169 motif that defines a superfamily of phosphotransferases. Our data suggest that the 3′-phosphatase active site incorporates multiple additional functional groups.  相似文献   

4.
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   

5.
Within the five classes (α, β, γ, δ, and ζ) of carbonic anhydrases (CAs) the first two, containing mammal and plant representatives, are the most studied among all CAs. In this study, we have focused our investigation on the beta-class carbonic anhydrase of Methanobacterium thermoautotrophicum. We investigated both the importance of the Asp-Arg dyad near the catalytic zinc-bound water and the possible roles that water molecules within the active site and residues near the entrance of the catalytic cleft have on the first step of the enzyme’s reaction mechanism. Hydrogen-bonding analysis of selected residues within the active site and constant pH replica exchange molecular dynamics constant pH replica exchange simulations were performed. The latter was done in order to evaluate the pKa values of possible proton acceptors. We found an intricate hydrogen-bonding network involving two acidic residues within the active site, Asp16 and Asp34, and the catalytic water molecule. We also observed a very strong interaction between the zinc-bound water and residues Asp34 and Arg36. This interaction was not significantly affected by the change in the protonation state of both the catalytic water and aspartate residue 34. The pKa analysis show that the effect of the R36A mutation affects not only the possible proton acceptors, but also the catalytic water itself.  相似文献   

6.
Studied was the effect of temperature in the range 12–46 °C on the rate of bacterial decolorization of the mono-azo dye Acid Orange 7 by Alcaligenes faecalis 6132 and Rhodococcus erythropolis 24. With both strains the raise of temperature led to a corresponding raise of decolorization rate better manifested by R. erythropolis. The analysis of the Arrhenius plot revealed a break near the middle of the temperature range. The regression analysis showed practically complete identity of the observed break point temperatures (T BP): 20.7 °C for Alc. faecalis and 20.8 °C for R. erythropolis. The values of the activation energy of the decolorization reaction (E a) were found to depend on both the organism and the temperature range. In the range below T BP the estimated values of E a were 138 ± 7 kJ mol−1 for Alc. faecalis and 160 ± 8 kJ mol−1 for R. erythropolis. In the range above T BP they were 54.2 ± 1.8 kJ mol−1 for Alc. faecalis and 37.6 ± 4.1 kJ mol−1 for R. erythropolis. Discussed are the possible reasons for the observed abrupt change of the activation energy.  相似文献   

7.
Molecularly imprinted polymeric membranes with tetrapeptide residue H-Asp(OcHex)-Asp(OcHex)-Asp(OcHex)-Asp(OcHex)-CH2- (DDDD) or H-Glu(OBzl)-Glu(OBzl)-Glu(OBzl)-Glu(OBzl)-C H2- (EEEE) were prepared during membrane preparation (casting) processing in the presence of print molecules. The Boc-L-Trp imprinted polymeric membranes thus obtained showed adsorption selectivity toward Ac-L-Trp from its racemic mixtures. From adsorption isotherms of Ac-Trp, the chiral recognition site, that had been formed by the presence of print molecules in the membrane preparation process, exclusively recognized Ac-L-Trp that possessed the same configuration of the print molecule. The affinity constants between chiral recognition sites in the membrane and Ac-L-Trp was determined to be 1.00 × 104 mol–1 dm3 and 1.08 × 104 mol–1 dm3 for the DDDD and EEEE membranes, respectively. Enantioselective electrodialysis could be attained by applying an optimum potential difference to give permselectivity, with a value close to its adsorption selectivity.  相似文献   

8.
Isothermal titration calorimetry (ITC) is shown to be a sensitive reporter of bile salt micellization and chiral recognition. Detailed ITC characterization of bile micelle formation as well as the chiral recognition capabilities of sodium cholate (NaC), deoxycholate (NaDC), and taurodeoxycholate (NaTDC) micelle systems are reported. The ΔHdemic of these bile salt micelle systems is directly observable and is strongly temperature‐dependent, allowing also for the determination of ΔCpdemic. Using the pseudo‐phase separation model, ΔGdemic and TΔSdemic were also calculated. Chirally selective guest–host binding of model racemic compounds 1,1’‐bi‐2‐napthol (BN) and 1,1’‐binaphthyl‐2,2’‐diylhydrogenphosphate (BNDHP) to bile salt micelles was then investigated. The S‐isomer was shown to bind more tightly to the bile salt micelles in all cases. A model was developed that allows for the quantitative determination of the enthalpic difference in binding affinity that corresponds to chiral selectivity, which is on the order of 1 kJ mol‐1. Chirality 28:290–298, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

9.
The enthalpy change for phosphorylation of ADP3? by PEP3? catalysed by pyruvate kinase has been determined at 25°C using flow microcalorimetry. Measurements were made at pH 8 in three buffer systems TRIS, TEA and HEPES and also at pH 8.5 in TRIS buffer. The values of ΔH obtained, ?8.75 kJ mol?1 in TRIS, ?7.39 kJ mol? in TEA and ?6.19 kJ mol?1 in HEPES surprisingly display a dependence on the buffer system used. The enthalpy change was combined with free energy data to calculate the entropy change for the catalysed reaction.  相似文献   

10.
Residue-specific chemical modification of amino acid residues of the microsomal epoxide hydrolase (mEH) from Rhodosporidium toruloides UOFS Y-0471 revealed that the enzyme is inactivated through modification of Asp/Glu and His residues, as well as through modification of Ser. Since Asp acts as the nucleophile, and Asp/Glu and His serve as charge relay partners in the catalytic triad of microsomal and soluble epoxide hydrolases during epoxide hydrolysis, inactivation of the enzyme by modification of the Asp/Glu and His residues agrees with the established reaction mechanism of these enzymes. However, the inactivation of the enzyme through modification of Ser residues is unexpected, suggesting that a Ser in the catalytic site is indispensable for substrate binding by analogy of the role of Ser residues in the related L-2-haloacid dehalogenases, as well as the ATPase and phosphatase enzymes. Co2+, Hg2+, Ag+, Mg2+ and Ca2+ inhibited enzyme activity and EDTA increased enzyme activity. The activation energy for inactivation of the enzyme was 167 kJ mol–1. Kinetic constants for the enzyme could not be determined since unusual behaviour was displayed during hydrolysis of 1,2-epoxyoctane by the purified enzyme. Enantioselectivity w as strongly dependent on substrate concentration. When the substrate was added in concentrations ensuring two-phase conditions, the enantioselectivity was greatly enhanced. On the basis of these results, it is proposed that this enzyme acts at an interface, analogous to lipases.  相似文献   

11.
Melanocortin-4 receptor (MC4R) has an important regulatory role in energy homeostasis and food intake. Peptide agonists of the MC4R are characterized by the conserved sequence His6-Phe7-Arg8-Trp9, which is crucial for their interaction with the receptor. This investigation utilized the covalent attachment approach to identify receptor residues in close proximity to the bound ligand [Nle4,d-Phe7]melanocyte-stimulating hormone (NDP-MSH), thereby differentiating between residues directly involved in ligand binding and those mutations that compromise ligand binding by inducing conformational changes in the receptor. Also, recent X-ray structures of G-protein-coupled receptors were utilized to refine a model of human MC4R in the active state (R?), which was used to generate a better understanding of the binding mode of the ligand NDP-MSH at the atomic level.The mutation of residues in the human MC4R—such as Leu106 of extracellular loop 1, and Asp122, Ile125, and Asp126 of transmembrane (TM) helix 3, His264 (TM6), and Met292 (TM7)—to Cys residues produced definitive indications of proximity to the side chains of residues in the core region of the peptide ligand. Of particular interest was the contact between d-Phe7 on the ligand and Ile125 of TM3 on the MC4R. Additionally, Met292 (TM7) equivalent to Lys(7.45) (Ballesteros numbering scheme) involved in covalently attaching retinal in rhodopsin is shown to be in close proximity to Trp9.For the first time, the interactions between the terminal regions of NDP-MSH and the receptor are described. The amino-terminus appears to be adjacent to a series of hydrophilic residues with novel interactions at Cys196 (TM5) and Asp189 (extracellular loop 2). These interactions are reminiscent of sequential ligand binding exhibited by the β2-adrenergic receptor, with the former interaction being equivalent to the known interaction involving Ser204 of the β2-adrenergic receptor.  相似文献   

12.
Cysteine synthesis in bacteria and plants is catalyzed by serine acetyltransferase (SAT) and O-acetylserine (thiol)-lyase (OAS-TL), which form the hetero-oligomeric cysteine synthase complex (CSC). In plants, but not in bacteria, the CSC is assumed to control cellular sulfur homeostasis by reversible association of the subunits. Application of size exclusion chromatography, analytical ultracentrifugation, and isothermal titration calorimetry revealed a hexameric structure of mitochondrial SAT from Arabidopsis thaliana (AtSATm) and a 2:1 ratio of the OAS-TL dimer to the SAT hexamer in the CSC. Comparable results were obtained for the composition of the cytosolic SAT from A. thaliana (AtSATc) and the cytosolic SAT from Glycine max (Glyma16g03080, GmSATc) and their corresponding CSCs. The hexameric SAT structure is also supported by the calculated binding energies between SAT trimers. The interaction sites of dimers of AtSATm trimers are identified using peptide arrays. A negative Gibbs free energy (ΔG = −33 kcal mol−1) explains the spontaneous formation of the AtCSCs, whereas the measured SAT:OAS-TL affinity (KD = 30 nm) is 10 times weaker than that of bacterial CSCs. Free SAT from bacteria is >100-fold more sensitive to feedback inhibition by cysteine than AtSATm/c. The sensitivity of plant SATs to cysteine is further decreased by CSC formation, whereas the feedback inhibition of bacterial SAT by cysteine is not affected by CSC formation. The data demonstrate highly similar quaternary structures of the CSCs from bacteria and plants but emphasize differences with respect to the affinity of CSC formation (KD) and the regulation of cysteine sensitivity of SAT within the CSC.  相似文献   

13.
A loop (residues 794–803) at the active site of β-galactosidase (Escherichia coli) opens and closes during catalysis. The α and β carbons of Ser-796 form a hydrophobic connection to Phe-601 when the loop is closed while a connection via two H-bonds with the Ser hydroxyl occurs with the loop open. β-Galactosidases with substitutions for Ser-796 were investigated. Replacement by Ala strongly stabilizes the closed conformation because of greater hydrophobicity and loss of H-bonding ability while replacement with Thr stabilizes the open form through hydrophobic interactions with its methyl group. Upon substitution with Asp much of the defined loop structure is lost. The different open-closed equilibria cause differences in the stabilities of the enzyme · substrate and enzyme · transition state complexes and of the covalent intermediate that affect the activation thermodynamics. With Ala, large changes of both the galactosylation (k2) and degalactosylation (k3) rates occur. With Thr and Asp, the k2 and k3 were not changed as much but large ΔH3 and TΔS3 changes showed that the substitutions caused mechanistic changes. Overall, the hydrophobic and H-bonding properties of Ser-796 result in interactions strong enough to stabilize the open or closed conformations of the loop but weak enough to allow loop movement during the reaction.  相似文献   

14.
A novel series of imidazole‐linked thiazolidinone hybrid molecules were designed and synthesized through a feasible synthetic protocol. The molecules were characterized with Fourier transform infrared (FT‐IR), 1H nuclear magnetic resonance (NMR), 13C NMR and high‐resolution mass spectrometry (HRMS) techniques. In vitro susceptibility tests against Gram‐positive (S. aureus and B. subtilis ) and Gram‐negative bacteria (E. coli and P. aeruginosa ) gave highly promising results. The most active molecule (3e) gave a minimal inhibitory concentration (MIC) value of 3.125 μg/mL which is on par with the reference drug streptomycin. Structure–activity relationships revealed activity enhancement by nitro and chloro groups when they occupied meta position of the arylidene ring in 2‐((3‐(imidazol‐1‐yl)propyl)amino)‐5‐benzylidenethiazolidin‐4‐ones. DNA‐binding study of the most potent molecule 3e with salmon milt DNA (sm‐DNA) under simulated physiological pH was probed with UV–visible absorption, fluorescence quenching, gel electrophoresis and molecular docking techniques. These studies established that compound 3e has a strong affinity towards DNA and binds at DNA minor groove with a binding constant (Kb) 0.18 × 102 L mol?1. Molecular docking simulations predicted strong affinity of 3e towards DNA with a binding affinity (ΔG) ‐8.5 kcal/mol. Van der Waals forces, hydrogen bonding and hydrophobic interactions were predicted as the main forces of interaction. The molecule 3e exhibited specific affinity towards adenine–thiamine base pairs. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

15.
Using directed evolution and site‐directed mutagenesis, we have isolated a highly thermostable variant of Aspergillus niger glucoamylase (GA), designated CR2‐1 . CR2‐1 includes the previously described mutations Asn20Cys and Ala27Cys (forming a new disulfide bond), Ser30Pro, Thr62Ala, Ser119Pro, Gly137Ala, Thr290Ala, His391Tyr and Ser436Pro. In addition, CR2‐1 includes several new putative thermostable mutations, Val59Ala, Val88Ile, Ser211Pro, Asp293Ala, Thr390Ser, Tyr402Phe and Glu408Lys, identified by directed evolution. CR2‐1 GA has a catalytic efficiency (kcat/Km) at 35°C and a specific activity at 50°C similar to that of wild‐type GA. Irreversible inactivation tests indicated that CR2‐1 increases the free energy of thermoinactivation at 80°C by 10 kJ mol?1 compared with that of wild‐type GA. Thus, CR2‐1 is more thermostable (by 5 kJ mol?1 at 80°C) than the most thermostable A. niger GA variant previously described, THS8 . In addition, Val59Ala and Glu408Lys were shown to individually increase the thermostability in GA variants by 1 and 2 kJ mol?1, respectively, at 80°C.  相似文献   

16.
The interaction of fisetholz with bovine serum albumin (BSA) and human serum albumin (HSA) was investigated by multi-spectroscopic, cyclic voltammetric, and molecular docking technique. The results revealed that there was a static quenching of BSA/HSA induced by fisetholz. The binding constants (Ka) and binding sites (n) were calculated at different temperatures (293, 303, and 311?K). The enthalpy change (ΔH) were calculated to be –17.20?kJ mol?1 (BSA) and –18.28?kJ mol?1 (HSA) and the entropy change (ΔS) were calculated to be 35.41?J mol?1 (BSA) and 24.02?J mol?1 (HSA), respectively, which indicated that the interaction between fisetholz and BSA/HSA was mainly by electrostatic attraction. Based on displacement experiments using site probes, indomethacin and ibuprofen, the binding site of fisetholz to BSA/HSA was identified as sub-domain IIIA, which was further confirmed by molecular docking method. There was little effect of K+, Ca2+, Cu2+, Zn2+, and Fe3+ on fisetholz-BSA or fisetholz-HSA complex. The spectra of synchronous fluorescence, circular dichroism (CD) and Fourier transform infrared (FT-IR) all showed that fisetholz binding to BSA/HSA leads to secondary structures change of the two serum albumins. According to the Förster non-radiation energy transfer theory, the binding distance between fisetholz and BSA/HSA was 2.94/4.68?nm. The cyclic voltammetry as a supporting tool also indicated that fisetholz interacted with protein.

Communicated by Ramaswamy H. Sarma  相似文献   


17.
A profound influence of water has previously been detected in the complexation of the enantiomers of methyl 2‐chloropropanoate (MCP) and the chiral selector octakis(3‐O‐butanoyl‐2,6‐di‐O‐pentyl)‐γ‐cyclodextrin (Lipodex‐E) in NMR and sensor experiments. We therefore investigated the retention behavior of MCP enantiomers on Lipodex‐E by gas chromatography (GC) under hydrous conditions. Addition of water to the N2 carrier gas modestly reduced the retention factors k of the enantiomers, notably for the second eluted enantiomer (S)‐MCP. This resulted in an overall decrease of enantioselectivity ‐ΔS,R(ΔG) in the presence of water. The effect was fully reversible. Consequently, for a conditioned column in the absence of residual water, the determined thermodynamic data, i.e. ΔS,R(ΔH) = –12.64 ± 0.08 kJ mol‐1 and ΔS,R(ΔS) = –28.18 ± 0.23 J K‐1 mol‐1, refer to a true 1:1 complexation process devoid of hydrophobic hydration. Chirality 28:124–131, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

18.
Some properties of purified endo-l,4-β-D-xylanase (1,4-β-D-xylan xylanohydrolase, EC 3.2.1.8) from the ligniperdous fungusTrametes hirsuta were investigated. The enzyme was stable between pH 4.0 and 8.0 with optimum activity at pH 5.0–5.5. The temperature optimum was 50 °C and the enzyme was stable for up to 30 min at 45 °C; however, it was denatured at higher temperatures. TheK m for 4-O-methylgluourono-D-xylan was 6.36. 10−3 equivalents ofD-xylose per litre, the activation energy was 28 kJ mol−1. The molecular weight determined by means of gel chromatography was 22000–24000. The enzyme was activated by Ca2+ and inhibited by Ag+ and Hg2+. On the basis of the effect of 2-hy-droxy-5 nitrobenzyl bromide, N-bromosuccimmide and N-aeetyhmidazole it may be assumed that trytophan and possibly tyrosine residues influence the enzyme catalysis.  相似文献   

19.
Abstract The uptake of ammonium and nitrate nitrogen by cultured plants of the green freshwater alga Chara hispida L. has been compared quantitatively with the contribution of its rhizoidal tissue. In the short-term, the rhizoid takes up 7–20% of the ammonium nitrogen, and about 15% of the nitrate that is taken up by whole plants under similar conditions. The uptake was studied over a range of both temperatures and external concentrations. The apparent activation energy for the uptake of NH4+ and NO3? by the whole plant was found to be 50 kJ mol?1 and 30 kJ mol?1, respectively. For the rhizoid, the values were similar for both nitrogenous ions, 106 kJ mol?1 and 70–100 kJ mol?1. The rhizoidal uptake mechanism for ammonium nitrogen operates more efficiently compared to that in the whole plant. Nitrate is taken up by the rhizoid by a mechanism with a substrate affinity higher than in the plant taken as a whole. The possible ecological significance of the results is discussed.  相似文献   

20.
Calorimetric studies of the reduction of free oxygen in solution by sodium dithionite are in agreement with a stoichiometry of 2 moles Na2S2O4 per mole of oxygen. The reaction is biphasic with ΔHt - 118±7 kcal mol?1 (?494 ± 29 kJ mol?1). The initial phase of the reaction proceeds with an enthalpy change of ca ?20 kcal (?84 kJ) and occurs when 0.5 moles of dithionite have been added per mole dioxygen present. This could be interpreted as the enthalpy change for the addition of a single electron to form the superoxide anion. Further reduction of the oxygen to water by one or more additional steps is accompanied by an enthalpy change of ca ?100 kcal (?418. 5 kJ). Neither of these reductive phases is consistent with the formation of hydrogen peroxide as an intermediate. The reduction of hydrogen peroxide by dithionite in 0.1 M phosphate buffer, pH 7.15, is a much slower process and with an enthalpy change of ca ? 74 kcal mol?1 (?314 kJ mol?1). Dissociation of oxyhemoglobin induced by the reduction of free oxygen tension with dithionite also shows a stoichiometry of 2 moles dithionite per mole oxygen present and an enthalpy change of ca. ?101 ±9 kcal mol?1 (?423± 38 kJ mol?1). The difference in the observed enthalpies (reduction of dioxygen vs. oxyhemoglobin) has been attributed to the dissociation of oxyhemoglobin, which is 17 kcal mol?1 (71 kJ mol?1).  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