首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Studies of abnormal and chemically modified haemoglobins indicate that in 0.1 m-NaCl about 40% of the alkaline Bohr effect of human haemoglobin is contributed by the C-terminal histidine HC3(146)β. In deoxyhaemoglobin, the imidazole of this histidine forms a salt bridge with aspartate FG1(94)β, in oxyhaemoglobin or carbonmonoxyhaemoglobin it accepts a hydrogen bond from its own NH group instead. Kilmartin et al. (1973) showed that in 0.2 m-NaCl + 0.2 m-phosphate this change of ligation lowered the pKa of the histidine from 8.0 in Hb3 to 7.1 in HbCO, but Russu et al. (1980) claimed that in bis-Tris buffer without added NaCl its pKa in HbCO dropped no lower than 7.85, and that in this medium the C-terminal histidine made only a negligible contribution to the alkaline Bohr effect.We have compared the histidine resonances of HbCO A with those of three abnormal haemoglobins: HbCO Cowtown (His HC3(146)β → Leu), HbCO Wood (His FG4(97)β → Leu) and HbCO Malmø (His FG4(97)β → Gln). Our results show that the resonance assigned by Russu et al. to His HC3(146)β in fact belongs to His FG4(97)β. Although in Hb the pKa of His HC3(146)β is 8.05 ± 0.05 independent of ionic strength, in HbCO its pKa drops sharply with diminishing ionic strength, so that in the buffer employed by Russu et al. it has a pKa of 6.2 and makes a contribution to the alkaline Bohr effect that is 57% larger than in the phosphate buffer employed by Kilmartin et al. (1973).In HbCO A, His FG4(97)β does not contribute to the Bohr effect, but in HbCO from which His HC3(146)β has been cleaved (HbCO des-His), His FG4(97)β is in equilibrium between two conformations with different pKa values. This equilibrium varies with ionic strength and pH, and presumably also with degree of ligation of the haem moiety.In HbCO A, His FG4(97)β has a pKa of 7.8 compared to the pKa value of about 6.6 characteristic of free histidines at the surface of proteins. This high pKa is accounted for by its interaction with the negative pole at the C terminus of helices F and FG. It corresponds to a free energy change of the same order as that observed in the interaction of histidines with carboxylate ions and confirms the strongly dipolar character of α-helices, which manifests itself even when they lie on the surface of the protein.  相似文献   

2.
Crystals of deoxyhaemoglobin Yakima (Asp Gl(99)β → His) are isomorphous with those of deoxyhaemoglobin A, even though the mutation produces disturbances in both the tertiary structure of the subunits and the quaternary structure of the tetramer. Asp Gl(99)β2 lies at the α1β2 subunit interface, and in deoxyhaemoglobin A forms a crucial hydrogen bond with Tyr C7(42) α1. The histidine residue that replaces the aspartate results in the removal of this single important intersubunit bond, and it further acts as a wedge between the α1 and β2 subunits, so that they are pushed apart and displaced part of the way towards the oxy structure. These disturbances are accompanied by the formation of a new intersubunit hydrogen bond, which is usually only observed in the oxy quaternary structure of haemoglobin. The disturbances at the α1β2 contact affect the stereochemistry of the entire molecule and are transmitted to the α and β haems. The X-ray structure of deoxy Yakima therefore provides a stereochemical explanation for its abnormal function; this being an abnormally high affinity for oxygen and vastly diminished haem-haem interactions.  相似文献   

3.
We have purified haemoglobin Philly by isoelectric focusing on polyacrylamide gel, and studied its oxygen equilibrium, proton nuclear magnetic resonance spectra, mechanical stability, and pH-dependent u.v. difference spectrum. Stripped haemoglobin Philly binds oxygen non-co-operatively with high affinity. Inorganic phosphate and 2,3-diphosphoglycerate have little effect on the equilibrium curve, but inositol hexaphosphate lowers the affinity and induces co-operativity. These properties are explained by the nuclear magnetic resonance spectra which show that stripped deoxyhaemoglobin Philly has the quaternary oxy structure and that inositol hexaphosphate converts it to the deoxy structure. An exchangeable proton resonance at ?8.3 p.p.m. from water, which is present in oxy- and deoxyhaemoglobin A, is absent in both these derivatives of haemoglobin Philly and can therefore be assigned to one of the hydrogen bonds made by tyrosine C1-(35)β, probably the one to aspartate H8(126)α at the α1β1 contact. Haemoglobin Philly shows the same pH-dependent u.v. difference spectrum as haemoglobin A, only weaker, so that a tyrosine other than 35β must be mainly responsible for this.  相似文献   

4.
Des arginine 141 a haemoglobin (the haemoglobin in which the C-terminal arginine of the a chain has been removed) has a high affinity for oxygen and a reduced co-operativity in its oxygen equilibrium binding. The kinetic consequences of this modification are investigated in this paper. Deoxy des Arg haemoglobin binds carbon monoxide faster than does haemoglobin A, whilst oxy des Arg haemoglobin loses oxygen more slowly. These results are correlated with the oxygen equilibrium binding properties of des Arg haemoglobin. The carbon monoxide binding kinetics have been interpreted as implying a change in the parameter c (of the allosteric model), as well as L, when this arginine is removed from haemoglobin.  相似文献   

5.
The binding of NADH to the dimeric (αβ) and tetrameric (α2β2) states of human aquomethaemoglobin has been characterized by sedimentation equilibrium studies of the effect of the concentration of free ligand on the macromolecular state of the haemoprotein. Both macromolecular states of aquomethaemoglobin exhibit a single binding site for NADH, which interacts approximately tenfold more strongly (6000 cf. 700 M−1) with the tetramer under the conditions studied (pH 6.0, I 0.10). Because the structure of aquomethaemoglobin resembles that of the deoxy state of haemoglobin, there is a high probability that organic phosphates also bind to dimeric deoxyhaemoglobin, a phenomenon which is not considered in thermodynamic treatments of the interplay between oxygen binding and its allosteric inhibition by 2,3-bisphosphoglycerate. Fortunately, the equilibrium constant for deoxyhaemoglobin self-association is so large that neglect of the interaction between allosteric inhibitor and dimeric haemoglobin is an oversight that should have no deleterious implications in the resultant thermodynamic analysis of the interplay between the preferential interactions of oxygen and organic phosphate with the various macromolecular states of deoxyhaemoglobin.  相似文献   

6.
A re-examination of the C-2 histidine proton resonances of haemoglobins A and Cowtown (His HC3(146)β → Leu) in chloride-free Hepes buffer has shown that all the resonances present in haemoglobin A are present in haemoglobin Cowtown, so that the pKa of His HC3(146)β cannot be determined by nuclear magnetic resonance in this buffer.  相似文献   

7.
Alzheimer's disease is a progressive neurodegenerative disease characterized by extracellular deposits of β‐amyloid (Aβ) plaques. Aggregation of the Aβ42 peptide leading to plaque formation is believed to play a central role in Alzheimer's disease pathogenesis. Anti‐Aβ monoclonal antibodies can reduce amyloid plaques and could possibly be used for immunotherapy. We have developed a monoclonal antibody C706, which recognizes the human Aβ peptide. Here we report the crystal structure of the antibody Fab fragment at 1.7 Å resolution. The structure was determined in two crystal forms, P21 and C2. Although the Fab was crystallized in the presence of Aβ16, no peptide was observed in the crystals. The antigen‐binding site is blocked by the hexahistidine tag of another Fab molecule in both crystal forms. The poly‐His peptide in an extended conformation occupies a crevice between the light and heavy chains of the variable domain. Two consecutive histidines (His4–His5) stack against tryptophan residues in the central pocket of the antigen‐binding surface. In addition, they form hydrogen bonds to the acidic residues at the bottom of the pocket. The mode of his‐tag binding by C706 resembles the Aβ recognition by antibodies PFA1 and WO2. All three antibodies recognize the same immunodominant B‐cell epitope of Aβ. By similarity, residues Phe–Arg–His of Aβ would be a major portion of the C706 epitope. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

8.
HAEMOGLOBIN Hiroshima is a variant with interesting physiological properties1,2 discovered in a Japanese family. Its Bohr effect is halved, its oxygen affinity at physiological pH increased about three-fold and haem-haem interaction is somewhat reduced compared with normal haemoglobin. In 0.1 M NaCl solutions initially stripped of phosphate, 2,3-diphospho-glycerate (2,3-DPG) diminishes the oxygen affinity as in haemoglobin A (H. F. Bunn, unpublished results). The amino-acid substitution originally deduced for this abnormal haemoglobin was histidine 143 (H21)β ? aspartic acid1. It was possible to conceive of a mechanism which accounted for its diminished Bohr effect3, but the normal response of its oxygen affinity to 2,3-DPG was inconsistent with the proposed role of histidine 143 in 2,3-DPG binding by haemoglobin A4,5. An X-ray crystallographic study of deoxyhaemoglobin Hiroshima has now revealed that the replacement occurs not in position 143 but 146β. This was confirmed by chemical methods and the physiological properties of this haemoglobin are now satisfactorily accounted for. The results support the role of histidine 146β in the alkaline Bohr effect6.  相似文献   

9.
Using complementary approaches of potentiometry and NMR spectroscopy, we have determined that the equilibrium acid dissociation constant (pKa value) of the arginine guanidinium group is 13.8 ± 0.1. This is substantially higher than that of ∼12 often used in structure-based electrostatics calculations and cited in biochemistry textbooks. The revised intrinsic pKa value helps explains why arginine side chains in proteins are always predominantly charged, even at pH values as great as 10. The high pKa value also reinforces the observation that arginine side chains are invariably protonated under physiological conditions of near neutral pH. This occurs even when the guanidinium moiety is buried in a hydrophobic micro-environment, such as that inside a protein or a lipid membrane, thought to be incompatible with the presence of a charged group.  相似文献   

10.
Haemoglobin Aalborg (Gly74 (E18)beta----Arg) has a reduced oxygen affinity, in both the absence and the presence of organic phosphates; it has a raised affinity for organic phosphates, and it is moderately unstable. By contrast, haemoglobin Shepherds Bush (Gly74 (E18)beta----Asp) has an increased oxygen affinity in both the absence and the presence of organic phosphates, a diminished affinity for organic phosphates and is also unstable. We have determined the crystal structure of deoxyhaemoglobin Aalborg at 2.8 A resolution and compared it to the structures of deoxy- and oxyhaemoglobin A and of deoxyhaemoglobin Shepherds Bush. The guanidinium group of Arg74(E18)beta protrudes from the haem pocket and donates hydrogen bonds to the E and F helices. The carboxylate group of Asp74(E18)beta forms a hydrogen bond only with residue EF6 and is partially buried, which may be why haemoglobin Shepherds Bush appears to be more unstable than haemoglobin Aalborg. To discover why the latter has a low oxygen affinity, we superimposed the B, G and H helices of haemoglobin A, whose conformation is known to be unaffected by ligand binding, on those of haemoglobin Aalborg. This also brought helices E and the haems into superposition, but revealed a shift of the F helix of deoxyhaemoglobin Aalborg towards the EF-corner. This shift is opposite to that which occurs on ligand binding and on transition to the quaternary oxy-structure, and is linked to an increased tilt of the proximal histidine residue away from the haem axis. Since the relative positions of helices E and F and of the haem group are thought to be the main determinants of the changes in oxygen affinity, the shift of helix F may account for the reduced oxygen affinity of haemoglobin Aalborg. The shift may be due to a combination of steric and electrostatic effects introduced by the arginine residue's side-chain. The effects of the arginine and aspartate substitutions at position E18 beta on the 2,3-diphosphoglycerate affinity are equal and opposite. They can be quantitatively accounted for by the electrostatic attraction or repulsion by the oppositely charged side-chains.  相似文献   

11.
Amyloid-β (Aβ) peptides are implicated in the neurodegeneration of Alzheimer’s disease (AD). We previously investigated the mechanism of neurotoxicity of Aβ and found that human Aβ (huAβ) binds and depletes heme, forming an Aβ-heme complex with peroxidase activity. Rodent Aβ (roAβ) is identical to huAβ, except for three amino acids within the proposed heme-binding motif (Site-H). We studied and compared heme-binding between roAβ and huAβ. Unlike roAβ, huAβ binds heme tightly (Kd = 140 ± 60 nM) and forms a peroxidase. The plot of bound (huAβ-heme) vs. unbound heme fits best to a two site binding hyperbola, suggesting huAβ possesses two heme-binding sites. Consistently, a second high affinity heme-binding site was identified in the lipophilic region (site-L) of huAβ (Kd = 210 ± 80nM). The plot of (roAβ-heme) vs. unbound heme, on the other hand, was different as it fits best to a sigmoidal binding curve, indicating different binding and lower affinity of roAβ for heme (Kd = 1 μM). The effect of heme-binding to site-H on heme-binding to site-L in roAβ and huAβ is discussed. While both roAβ and huAβ form aggregates equally, rodents lack AD-like neuropathology. High huAβ/heme ratio increases the peroxidase activity. These findings suggest that depletion of regulatory heme and formation of Aβ-heme peroxidase contribute to huAβ’s neurotoxicity in the early stages of AD. Phylogenic variations in the amino acid sequence of Aβ explain tight heme-binding to huAβ and likely contribute to the increased human susceptibility to AD.  相似文献   

12.
A study was conducted to assess the effect of feeding graded levels of tannin-containing Prosopis cineraria leaves in a complete feed mixture (CFM) on the performance of lambs and kids. Eighteen lambs and 18 kids of Malpura and Marwari breed, respectively of similar age (90 days) and body weight (11.0 kg), were randomly distributed in to three groups of six each. Each group was offered complete feed mixtures ad libitum in feeding troughs under group feeding system for 90 days. The concentrate component of CFM was from the commercial feed in mash form containing corn, soybean meal, wheat bran and de-oiled rice bran. Prosopis leaves were ground to pass through a 4 mm sieve before being thoroughly mixed with required quantity of concentrate mash. CFMs contained graded levels of P. cineraria leaves. In, CFM-1, ratio of P. cineraria to concentrate mixture was 25: 75 (T1), whereas in CFM-2 and CFM-3, it was 50:50 (T2) and 75:25 (T3), respectively. The lambs in L1, L2 and L3 and kids in K1, K2 and K3, groups received CFM-1, CFM-2 and CFM-3, respectively. CP (g kg−1) was 182 in CFM-1, 162 in CFM-2 and 140 in CFM-3. P. cineraria leaves contained (g kg−1) CP 159, neutral detergent fibre (NDF) 567, acid detergent fibre (ADF) 360 and acid detergent lignin (ADL) 189 on dry matter (DM) basis. The extractable condensed tannin (CT; leucocyanidin equivalent), hydrolysable tannin (HT) and protein precipitation capacity (PPC) of the leaves (g kg−1) were 90.7, 3.4 and 111.5, respectively. DMI in L1 and L2 lambs were higher than their corresponding groups in kids, whereas it was reverse in L3 (more in kids than lambs). Significant difference in digestible crude protein (DCP) intake was recorded among the groups in both lambs and kids. In lambs, highest DCPI (g day−1) was recorded in L1 (93.3), as compared to L2 (82.2) and L3 (37.2), whereas in kids it was K2 (73.4), followed by K1 (69.2) and K3 (43.8). Significant difference in the digestibility of DM, CP and NDF were recorded among three groups in both species. Maximum nutrient digestibility was recorded in T1 followed by T2 and T3 in both lambs and kids. However, digestibility of all the nutrients was more in kids than that of lambs. Maximum ammonia-N, total N and trichloro acetic acid (TCA) precipitable N was recorded in T1 followed by T2 and T3 in both the species. However, there was no typical trend between the two species in rumen parameters. Blood haemoglobin (Hb) and blood urea nitrogen (BUN) values were highest in T1 followed by T2 and T3 in both lambs and kids. After 90 days of intensive feeding, maximum weight gain was recorded in T2 as compared to that of T1 and T3 in both lambs and kids. Although L1 lambs performed better under high concentrate diet compared to K1 kids, weight gain in K2 and K3 kids were significantly higher compared to their counter part lambs. Similar was the trend in ADG also. It was concluded that performance of lambs and kids differed with the level of CT in their diet and kids performed better on high tannin diet as compared to that of lambs.  相似文献   

13.
The accumulation of arginine in the cerebrospinal fluid and brains of patients suffering from acute neurodegenerative diseases like Alzheimer’s disease, point to defects in the metabolic pathways involving this amino acids. The deposits of neurofibrillary tangles and senile plaques perhaps as a consequence of fibrillogenesis of β-amyloid peptides has also been shown to be a hallmark in the aetiology of certain neurodegenerative diseases. Peptidylarginine deiminase (PAD II) is an enzyme that uses arginine as a substrate and we now show that PAD II not only binds with the peptides Aβ1-40, Aβ22-35, Aβ17-28, Aβ25-35 and Aβ32-35 but assists in the proteolytic degradation of these peptides with the concomitant formation of insoluble fibrils. PAD was purified in 12.5% yield and 137 fold with a specific activity of 59 μmol min?1?mg?1 from bovine brain by chromatography on diethylaminoethyl (DEAE)-Sephacel. Characterisation of the enzyme gave a pH and temperature optima of 7.5°C and 68°C, respectively, and the enzyme lost 50% activity within 38 min at this temperature. Michaelis-Menten kinetics established a V max and K m of 1.57 μmol min?1?ml?1 and 1.35 mM, respectively, with N-benzoyl arginine ethyl ester as substrate. Kinetic analysis was used to measure the affinity (K i) of the amyloid peptides to PAD with values between 1.4 and 4.6 μM. The formation of Aβ fibrils was rate limiting involving an initial lag time of about 24 h that was dependent on the concentration of the amyloid peptides. Turbidity measurements at 400 nm, Congo Red assay and Thioflavin-T staining fluorescence were used to establish the aggregation kinetics of PAD-induced fibril formation.  相似文献   

14.
Phosphite dehydrogenase (PTDH) catalyzes the NAD+-dependent oxidation of phosphite to phosphate. This reaction requires the deprotonation of a water nucleophile for attack on phosphite. A crystal structure was recently solved that identified Arg301 as a potential base given its proximity and orientation to the substrates and a water molecule within the active site. Mutants of this residue showed its importance for efficient catalysis, with about a 100-fold loss in k cat and substantially increased K m,phosphite for the Ala mutant (R301A). The 2.35 Å resolution crystal structure of the R301A mutant with NAD+ bound shows that removal of the guanidine group renders the active site solvent exposed, suggesting the possibility of chemical rescue of activity. We show that the catalytic activity of this mutant is restored to near wild-type levels by the addition of exogenous guanidinium analogues; Brønsted analysis of the rates of chemical rescue suggests that protonation of the rescue reagent is complete in the transition state of the rate-limiting step. Kinetic isotope effects on the reaction in the presence of rescue agents show that hydride transfer remains at least partially rate-limiting, and inhibition experiments show that K i of sulfite with R301A is ∼400-fold increased compared to the parent enzyme, similar to the increase in K m for phosphite in this mutant. The results of our experiments indicate that Arg301 plays an important role in phosphite binding as well as catalysis, but that it is not likely to act as an active site base.  相似文献   

15.
  • 1.1. The binding of O2 to goldfish haemoglobin showed a strong pH dependence P50=5.5 mmHg; n = 2.4 at pH 8.0 and P50 = 170 mmHg; n = 1.0 at pH 5.5 such that the protein is only 50% saturated in a solution of air equilibrated buffer at pH 5.5.
  • 2.2. The binding of CO is cooperative at high pH (n = 2.8; L = 1000; KR = 0.1 μM; KT = 4 μM) and non-cooperative (n = 1) at pH 5.5.
  • 3.3. The rate of O2 dissociation is extremely fast and pH dependent; being 30 sec−1 at pH 8.0 and 400 sec−1 at pH 6.0 at 1°C. At 23°C the rate of this process is too fast to obtain accurate data using stopped-flow techniques.
  • 4.4. Partial photolysis of the oxyhaemoglobin species leads to homogeneous recombination kinetics at pH 8.0 with an associated rate constant of 4.7 × 107 M−1 sec−1. At pH < 7.5 the recombination process occurs in two steps. One rate is equal to that observed at pH 8.0. The slower process is favoured at low pH.
  • 5.5. Photolysis of the CO haemoglobin complex indicates that, at high pH, combination of CO with deoxyhaemoglobin is cooperative, whilst recombination with Hb(CO)3 is non-cooperative and occurs at a rate of 1.2 × 106 M−1 sec−1.
  • 6.6. At neutral pH recombination of CO with partially linganded haemoglobin occurs in a two-step process. The proportion contributed by each of these two steps in pH dependent.
  • 7.7. The functioning of this Root effect haemoglobin is discussed in terms of the two state-model of cooperativity in which the αβ chain heterogeneity is minimal
  相似文献   

16.
Human fluoromethaemoglobin with inositol hexaphosphate (IHP) in 0.05 m-phosphate buffer was crystallized by addition of polyethylene glycol (PEG). The crystals are isomorphous with those of deoxyhaemoglobin A without IHP grown in solutions containing PEG by Ward et al. (1975). The structure was investigated by means of a difference Fourier synthesis against deoxyhaemoglobin A based on X-ray data collected within a limiting sphere of 3.5 Å?1. The four subunits are arranged in the quaternary T structure and IHP is bound at the same site between the β chains as in deoxyhaemoglobin. In both the α and β haem regions the distance between the haem plane and the F helix is reduced in fluoromethaemoglobin relative to deoxyhaemoglobin and the iron atom is moved from the proximal towards the distal side of the plane, but the change, if any, in the distance between the iron and the Nε of the proximal histidine cannot be clearly established. The α Fe in fluoromethaemoglobin is either in the haem plane or up to 0.8 Å on the distal side, suggesting the possibility of rupture of the bond to the histidines Nε; it was not possible to estimate the position of the β iron. The main spectral changes associated with the reaction of fluoromethaemoglobin with IHP take place in less than 3 ms at room temperature.  相似文献   

17.
Deposition of beta‐amyloid (Aβ) is considered as an important early event in the pathogenesis of Alzheimer's Disease (AD), and reduction of Aβ levels by various therapeutic approaches is actively being pursued. A potentially non‐inflammatory approach to facilitate clearance and reduce toxicity is to hydrolyze Aβ at its α‐secretase site. We have previously identified a light chain fragment, mk18, with α‐secretase‐like catalytic activity, producing the 1–16 and 17–40 amino acid fragments of Aβ40 as primary products, although hydrolysis is also observed following other lysine and arginine residues. To improve the specific activity of the recombinant antibody by affinity maturation, we constructed a single chain variable fragment (scFv) library containing a randomized CDR3 heavy chain region. A biotinylated covalently reactive analog mimicking α‐secretase site cleavage was synthesized, immobilized on streptavidin beads, and used to select yeast surface expressed scFvs with increased specificity for Aβ. After two rounds of selection against the analog, yeast cells were individually screened for proteolytic activity towards an internally quenched fluorogenic substrate that contains the α‐secretase site of Aβ. From 750 clones screened, the two clones with the highest increase in proteolytic activity compared to the parent mk18 were selected for further study. Kinetic analyses using purified soluble scFvs showed a 3‐ and 6‐fold increase in catalytic activity (kcat/KM) toward the synthetic Aβ substrate compared to the original scFv primarily due to an expected decrease in KM rather than an increase in kcat. This affinity maturation strategy can be used to select for scFvs with increased catalytic specificity for Aβ. These proteolytic scFvs have potential therapeutic applications for AD by decreasing soluble Aβ levels in vivo. © 2009 American Institute of Chemical Engineers. Biotechnol. Prog., 2009  相似文献   

18.
Gusta LV 《Plant physiology》1975,56(5):707-709
The freezing of water in acclimated and nonacclimated cereals was studied using pulsed nuclear magnetic resonance spectroscopy. The quantity of unfreezable water per unit dry matter was not strongly dependent on the degree of cold acclimation. In contrast, the fraction of water frozen which was tolerated by nonacclimated winter cereals and by an acclimated spring wheat (Triticum aestivum L.) was less than in acclimated hardy cereals. The freezing curves had the following form:LT = L0ΔTm/T + KLT and L0 are liquid water per unit dry matter at T and 0 C, respectively. ΔTm is the melting point depression and K is the liquid water which does not freeze.  相似文献   

19.
Laccases (benzenediol oxygen oxidoreductases, EC 1.10.3.2) are used in many biotechnological processes, including removal of polyphenols in beverages, decolorizing and detoxifying effluents, drug analysis and bioremediation. In the present work, we have tried to increase thermal stability of laccase from Bacillus HR03 using site directed point mutations. Glu188 was substituted with 2 positive (Lys and Arg) and one hydrophobic (Ala) residues. All mutations showed improved thermal stability. Thermal activation of laccase was also increased after introducing the mutations. Remarkably, the Glu188Lys variant showed 3-fold higher thermal activation and higher T50 (5 °C) with respect to the native enzyme. Furthermore steady-state kcat and Km values were influenced despite the distance between the mutated position and the catalytic site. In Glu188Arg mutation, the kcat was improved 3-fold and Km reduced by 25%. Interestingly, all three variants showed higher stability against urea as a chemical denaturant. Structural analyses of the native and mutated variants were carried out using fluorescence and far-UV circular dichroism.  相似文献   

20.
The maximum slope of the plot, appearing in the paper of Watari & Isogai (1976), was derived algebraically as a function of allosteric constants c and αmor βm (= m), and the relation between L, c, and αmor βm, was also obtained, where L = ToRo, c = KRKT, αm = FmKR, βm = FmKT, Roand To are concentrations of unligated R and T states respectively, KRand KT are microscopic dissociation constants, and Fm is the ligand concentration at the maximum slope of the plot. When the maximum slope is increased by one, the value becomes Hill constant, n. Nomographs which enable easier estimation of allosteric constants, L and c, were constructed from the two given values, the maximum slope of the plot, n ? 1, and αmor βm, in the cases where the maximum number of ligands, N, was 2 and 4. In the nomograph, log c is plotted against log L2cN keeping the value of the maximum slope of the plot and that of αmor βm constant. These nomographs show that the representation is symmetrical in the cases of L2cN > 1 and L2cN < 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