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 260 毫秒
1.
The pK values of the titratable groups in ribonuclease Sa (RNase Sa) (pI=3.5), and a charge-reversed variant with five carboxyl to lysine substitutions, 5K RNase Sa (pI=10.2), have been determined by NMR at 20 degrees C in 0.1M NaCl. In RNase Sa, 18 pK values and in 5K, 11 pK values were measured. The carboxyl group of Asp33, which is buried and forms three intramolecular hydrogen bonds in RNase Sa, has the lowest pK (2.4), whereas Asp79, which is also buried but does not form hydrogen bonds, has the most elevated pK (7.4). These results highlight the importance of desolvation and charge-dipole interactions in perturbing pK values of buried groups. Alkaline titration revealed that the terminal amine of RNase Sa and all eight tyrosine residues have significantly increased pK values relative to model compounds.A primary objective in this study was to investigate the influence of charge-charge interactions on the pK values by comparing results from RNase Sa with those from the 5K variant. The solution structures of the two proteins are very similar as revealed by NMR and other spectroscopic data, with only small changes at the N terminus and in the alpha-helix. Consequently, the ionizable groups will have similar environments in the two variants and desolvation and charge-dipole interactions will have comparable effects on the pK values of both. Their pK differences, therefore, are expected to be chiefly due to the different charge-charge interactions. As anticipated from its higher net charge, all measured pK values in 5K RNase are lowered relative to wild-type RNase Sa, with the largest decrease being 2.2 pH units for Glu14. The pK differences (pK(Sa)-pK(5K)) calculated using a simple model based on Coulomb's Law and a dielectric constant of 45 agree well with the experimental values. This demonstrates that the pK differences between wild-type and 5K RNase Sa are mainly due to changes in the electrostatic interactions between the ionizable groups. pK values calculated using Coulomb's Law also showed a good correlation (R=0.83) with experimental values. The more complex model based on a finite-difference solution to the Poisson-Boltzmann equation, which considers desolvation and charge-dipole interactions in addition to charge-charge interactions, was also used to calculate pK values. Surprisingly, these values are more poorly correlated (R=0.65) with the values from experiment. Taken together, the results are evidence that charge-charge interactions are the chief perturbant of the pK values of ionizable groups on the protein surface, which is where the majority of the ionizable groups are positioned in proteins.  相似文献   

2.
The instability of aqueous solutions of poly(α-L -glutamic acid) (PGA) at low pH is due to two distinguishable phenomena: precipitation, favored above 40°C., and aggregation, favored below 20°C. The aggregated form of PGA can be isolated by gel permeation chromatography. Both aggregation and precipitation increase with decreasing pH, i.e., with decreasing ionization of the side chain carboxyl groups. Temperature-induced aggregation and disaggregation give rise to a reproducible hysteresis loop which can be followed by optical rotation, light scattering, sedimentation, viscosity, and chromatography. Hysteresis has been observed with different PGA samples, and in several aqueous buffered or unbuffered solvents and organic-aqueous solvent mixtures and in the pH range 4.1–4.5. Aggregation manifests itself as an increase in negative optical rotation in the visible and ultraviolet spectral range. The specific relation at 233 mμ is sensitive to aggregation and also reflects the hysteresis. Measurements of optical rotatory dispersion indicate that a0 reflects the hysteresis but b0 does not, the latter revealing only reversible changes with aggregation and disaggregation. The helix-coil equilibrium is apparently unperturbed by aggregation, as is the thermal stability of the helix structure. For fully aggregated PGA it is estimated that a0 increases by about 300 degrees, which suggests that a0 may be a sensitive parameter to measure aggregation in other systems. The rate of aggregation increases with decreasing temperature. The disaggregation, upon heating, is more rapid. However, kinetics measurements have not yet been done. The temperature M at which all aggregates are disrupted increases with decreasing pH, but is independent of total PGA concentration, at constant pH. No molecular weight dependence of M was detected in the range 20–100 × 103. The shape and size of the hysteresis loop depends upon pH and molecular weight, which is interpreted as a dependence on the extent of aggregation. One branch of the loop, representing the helix–coil transition of isolated molecules, is reversible, while the others, representing the formation and disruption of the aggregates, are not. The system exhibits both ascending and descending scanning curves, which are typical of a true hysteresis.  相似文献   

3.
T Azuma  N Sakato  H Fujio 《Biochemistry》1988,27(16):6116-6120
The interaction of M315 with 2,4-dinitrophenyl haptens was studied. 2,4-Dinitroaniline (DNP-NH2) showed maximum affinity to M315 at about pH 4. The pH dependence of the association constant of DNP-NH2 to M315 showed three transitions at pH 4.7, at pH 7.2, and below pH 9, respectively. Since the DNP-NH2 molecule has no charged group in this pH range, the transitions were explained in terms of amino acid residues with ionizable side chains in M315. Judging from the pK values and the effect of succinylation, these transitions were concluded to be related to ionizations of carboxyl, imidazole, and phenol groups, respectively. Measurement of the fluorescence of affinity-labeled M315 suggested that the transition at pH 4.7 reflected an equilibrium between two forms of M315 with different conformations of the combining site. The contribution of the amino acid sequence on the light (L) chain to the interaction with haptens was studied by use of antibodies (Abs) reconstituted from the heavy chain of M315 (H315) and either a homologous or a heterologous L chain. The reconstituted heterologous Ab (H315L952) showed similar pH dependence of binding to DNP-NH2 to that of the homologous Ab (H315L315). Moreover, the two Abs showed no appreciable difference in binding to DNP-haptens of different sizes. These results suggested that the difference in the amino acid sequences of L315 and L952, which originated by a somatic hypermutation, has little effect on the ligand binding.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
T Kesvatera  B J?nsson  A Telling  V T?ugu  H Vija  E Thulin  S Linse 《Biochemistry》2001,40(50):15334-15340
The binding of calcium ions by EF-hand proteins depends strongly on the electrostatic interactions between Ca(2+) ions and negatively charged residues of these proteins. We have investigated the pH dependence of the binding of Ca(2+) ions by calbindin D(9k). This protein offers a unique possibility for interpretation of such data since the pK(a) values of all ionizable groups are known. The binding is independent of pH between 7 and 9, where maximum calcium affinity is observed. An abrupt decrease in the binding affinity is observed at pH values below 7. This decrease is due to protonation of acidic groups, leading to modification of protein charges. The pH dependence of the product of the two macroscopic Ca(2+)-binding constants can be formally described by the involvement of two acidic groups with pK(a) = 6.6. Monte Carlo calculations show that the reduction of Ca(2+) binding is strictly determined by variable electrostatic interactions due to pH-dependent changes not only in the binding sites, but also of the overall charge of the protein.  相似文献   

5.
When azide ion reacts with methemoglobin in unbuffered solution the pH of the solution increases. This phenomenon is associated with increases in the pK values of heme-linked ionizable groups on the protein which give rise to an uptake of protons from solution. We have determined as a functional of pH the proton uptake, delta h+, on azide binding to methemoglobin at 20 degrees C. Data for methemoglobins A (human), guinea pig and pigeon are fitted to a theoretical expression based on the electrostatic effect of these sets of heme-linked ionizable groups on the binding of the ligand. From these fits the pK values of heme-linked ionizable groups are obtained for liganded and unliganded methemoglobins. In unliganded methemoglobin pK1, which is associated with carboxylic acid groups, ranges between 4.0 and 5.5 for the three methemoglobins; pK2, which is associated with histidines and terminal amino groups, ranges from 6.2 to 6.7. In liganded methemoglobin pK1 lies between 5.8 and 6.3 and pK2 varies from 8.1 to 8.5. The pH dependences of the apparent equilibrium constants for azide binding to the three methemoglobins at 20 degrees C are well accounted for with the pK values calculated from the variation of delta h+ with pH.  相似文献   

6.
The ammonia‐oxidizing archaeon Nitrosopumilus maritimus strain SCM1 (strain SCM1), a representative of the Thaumarchaeota archaeal phylum, can sustain high specific rates of ammonia oxidation at ammonia concentrations too low to sustain metabolism by ammonia‐oxidizing bacteria (AOB). One structural and biochemical difference between N. maritimus and AOB that might be related to the oligotrophic adaptation of strain SCM1 is the cell surface. A proteinaceous surface layer (S‐layer) comprises the outermost boundary of the strain SCM1 cell envelope, as opposed to the lipopolysaccharide coat of Gram‐negative AOB. In this work, we compared the surface reactivities of two archaea having an S‐layer (strain SCM1 and Sulfolobus acidocaldarius) with those of four representative AOB (Nitrosospira briensis, Nitrosomonas europaea, Nitrosolobus multiformis, and Nitrosococcus oceani) using potentiometric and calorimetric titrations to evaluate differences in proton‐ionizable surface sites. Strain SCM1 and Sacidocaldarius have a wider range of proton buffering (approximately pH 10–3.5) than the AOB (approximately pH 10–4), under the conditions investigated. Thermodynamic parameters describing proton‐ionizable sites (acidity constants, enthalpies, and entropies of protonation) are consistent with these archaea having proton‐ionizable amino acid side chains containing carboxyl, imidazole, thiol, hydroxyl, and amine functional groups. Phosphorous‐bearing acidic functional groups, which might also be present, could be masked by imidazole and thiol functional groups. Parameters for the AOB are consistent with surface structures containing anionic oxygen ligands (carboxyl‐ and phosphorous‐bearing acidic functional groups), thiols, and amines. In addition, our results showed that strain SCM1 has more reactive surface sites than the AOB and a high concentration of sites consistent with aspartic and/or glutamic acid. Because these alternative boundary layers mediate interaction with the local external environment, these data provide the basis for further comparisons of the thermodynamic behavior of surface reactivity toward essential nutrients.  相似文献   

7.
The N-acetyl-beta-D-hexosaminidase B of germinating yellow lupin seeds catalyzed the hydrolysis of both p-nitrophenyl-N-acetyl-beta-D-glucosaminide and -galactosaminide substrates. The investigation of the pH dependence of the kinetic parameters (Vmax and Vmax/Km) demonstrated that two common ionizable groups (probably two carboxyl groups) play an essential role in the catalysis. That is, the enzyme has a lysozyme-like splitting mechanism, and the possibility of an anchimeric assistance provided by the acetamido group seems to be negligible. The presence of a deprotonated carboxyl group near the glycosidic linkage was also supported by inhibition with 1-thio substrate analogues. On the other hand, some 2-acetamido-2-deoxyaldono(1----4)lactones proved to be effective inhibitors of the hexosaminidase with the exception of the D-arabinose derivative, which can be explained by high stereospecificity in the binding.  相似文献   

8.
The dependence on pH of the fluorescence of cholera toxin and its A and B subunits has been studied at 25 degrees C. The fluorescence intensity of cholera toxin is highly pH-dependent. In the pH range 7-9.5 it reaches a maximum corresponding to a quantum yield of 0.076. In the pH range 4-7 a strong increase in fluorescence intensity is observed (delta Q/Qmax = 0.64). Evaluation of the pH sensitivity of the fluorescence intensity of the A and B subunits reveals that the B subunit is mainly responsible for the observed pH effect (delta Q/Qmax for B subunit = 0.64). The intensity changes are paralleled by similar although less pronounced changes in the average fluorescence excited state life-time tau (delta tau/tau max = 0.33 for cholera toxin). Fluorimetric titration of the B subunit, which is related to the indole fluorescence of the lone Trp-88, reveals that the fluorescence intensity changes in the pH range 4-7 are due to reaction of two types of ionizable quencher displaying apparent pKa values of 4.4 and 6.2, respectively. It is suggested that the increase in fluorescence intensity with a midpoint at pH 6.2 is the result of deionization of the imidazolium side-chain of one or two out of the four histidine residues present in each beta-polypeptide chain, whereas a deionized carboxyl group is responsible for the quenching with midpoint at pH 4.4. Complex formation of cholera toxin or B subunit with the monosialoganglioside GM1 or the oligosaccharide moiety of GM1 (oligo-GM1) completely prevents the quenching by both quenchers. Addition of 6 M urea also eliminates the pH effect. The quenching is not the result of the dissociation of the B subunit into its constituent monomers. Upon fluorimetric titration of cholera toxin or B subunit above pH 9, a progressive drop in both fluorescence intensity and tau occurs. This decrease could be due to energy transfer from the indole moiety of Trp-88 to ionized tyrosines or by quenching through an unprotonated epsilon-amino group of lysine. Fluorimetric titration of the A subunit indicates that the tryptophan fluorescence is only moderately altered by ionizable groups displaying a pKa in the range 4 to 9. Activation of A subunit does not affect this lack of pH sensitivity. Above pH 9, however, a much more significant drop in the fluorescence intensity of activated A subunit occurs. The structural implications of the results are discussed.  相似文献   

9.
Trehalase (EC 3.2.1.28) of the bound type was purified as an electrophoretically homogeneous protein from adult honeybees by fractionation with ammonium sulfate, hydrophobic chromatography, and DEAE-Sepharose CL-6B, CM-Sepharose CL-6B, butyl-Toyopearl 650M, and p-aminophenyl beta-glucoside Sepharose 4B column chromatographies. The enzyme preparation was confirmed to be a monomeric protein containing 3.1% carbohydrate. The molecular weight was estimated to be approximately 69,000, and the optimum pH was 6.7. The Michaelis constant (Km) was 0.66 mM, and the molecular activity (k0) was 86.2 s(-1). The enzyme was an "inverting" type which produced beta-glucose from alpha, alpha-trehalose. Dependence of the V and Km values on pH gave values for the ionization constants, pKe1 and pKe2, of essential ionizable groups 1 and 2 of the free enzyme of 5.3 and 8.5, respectively. When the dielectric constant of the reaction mixture was decreased, pKe1, and pKe2 were shifted to higher values of + 0.2 and + 0.5 pH unit, respectively. The ionization heat (deltaH) of ionizable group 1 was estimated to be + 1.8 kcal/mol, and the deltaH value of group 2 was + 1.5 kcal/mol. These findings strongly support the notion that the essential ionizable groups of honeybee trehalase are two kinds of carboxyl groups, one being a dissociated type (-COO(-), ionizable group 1) and the other a protonated type (-COOH, ionizable group 2), although the pKe2 value is high.  相似文献   

10.
Hydrogen ion titration of an affinity-purified mannose/glucose-specific lectin from Cajanus cajan pulse was carried out at 30°C and ionic strength of 0.15 by a discontinuous method. The titration was reversible in the pH range 2–12.0. The numbers of different ionizable groups per 39,000 g of the lectin were 43 carboxyl groups (pKint = 3.93), 10 imidazole groups, 21 ε-amino groups, 12.8 phenoxyl groups (pKint = 10.0), and 5 guanidyl groups. Only seven tyrosine residues of the lectin were dissociated under native conditions. The remaining six tyrosines became available for titration upon denaturation of the lectin in 9 M urea.  相似文献   

11.
M Murata  K Nagayama  S Ohnishi 《Biochemistry》1987,26(13):4056-4062
The membrane fusion activity of melittin and its succinylated derivative was studied as a function of pH by the transfer of spin-labeled phosphatidylcholine as well as by internal content mixing and electron microscopy. The protonation process of the carboxyl groups introduced into melittin was studied by 13C NMR spectroscopy using derivative prepared with [1,4(-13)C]succinic anhydride. Melittin causes fusion of sonicated phosphatidylcholine vesicles in a wide range of pH. In marked contrast, melittin with all four amino groups succinylated induces fusion only at acidic pH lower than 5.2, with the maximum at pH 5.1. The fusion reactions are very rapid, reaching a saturation level within 1 min. The fusion efficiency depends on the peptide-to-phospholipid ratio in the reaction mixture. Trypsinized succinylated melittin, which has lost the four positively charged C-terminal residues, causes aggregation of vesicles at acidic pH but cannot induce fusion. The 13C NMR peaks for the carboxyl and carbonyl groups of succinylated melittin shifted to higher field as the pH was lowered. The pKa value of the four carboxyl groups was obtained as 5.19 and 4.83 in the presence and absence of vesicles, respectively. The pKa value in the presence of vesicles agrees quite well with the half-maximal pH for fusion of 5.15, indicating that the fusion activity is triggered by protonation of the carboxyl groups in the hydrophobic segment of the peptide. The higher shift of pKa value in the presence of vesicles can be due to stabilization of the protonated form by entrance into lipid bilayer hydrocarbon layer.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
A series of hydrogels were synthesized by graft cross-link copolymerization of sodium alginate (SA) and acrylic acid (AA) using N, N-methylene-bis-(acrylamide) as a cross-linker. By study of the swelling kinetics of the hydrogels in different buffer solutions, the overshooting effect was observed in acidic medium, namely the gels firstly swelled to a maximum value following by a gradual deswelling until the equilibrium. The phenomenon is interpreted as a cooperative physical cross-linking caused by the hydrogen bond formation between the carboxyl groups of the hydrogels in a hydrophobic environment. The hydrogen bond formation was further confirmed by FT-IR spectra. The dependence of overshooting effect on the pH of buffer solution was more noticeable in comparison with the composition of hydrogels, demonstrating that the cooperative physical cross-linking caused by the hydrogen bond formation is dominant. Whether or not the overshooting effect appears is not only relative to the pH of buffer solution, but also depends on the pKa of carboxyl groups on the network. The overshoot processes of the hydrogels under acidic medium at pH below the pKa follow a quantitative model proposed by Díez-Peńa et al., and the theoretical curves are in very good agreement with the experimental data. While in pH > pKa buffer solutions, the overshoot phenomenon does not appear arising from the repulsive interaction between the ionized carboxyl groups, the swelling processes follow Schott second-order rate equation.  相似文献   

13.
We have used crystallography and thermodynamic analysis to study nuclease variants I92E and I92K, in which an ionizable side-chain is placed in the hydrophobic core of nuclease. We find that the energetic cost of burying ionizable groups is rather modest. The X-ray determinations show water molecules solvating the buried glutamic acid under cryo conditions, but not at room temperature. The lysine side-chain does not appear solvated in either case. Guanidine hydrochloride (GnHCl) denaturation of I92E and I92K, done as a function of pH and monitored by tryptophan fluorescence, showed that I92E and I92K are folded in the pH range pH 3.5-9.0 and pH 5.5-9.5, respectively. The stability of the parental protein is independent of pH over a broad range. In contrast, the stabilities of I92E and I92K exhibit a pH dependence, which is quantitatively explained by thermodynamic analysis: the PK(a) value of the buried K92 is 5.6, while that of the buried E92 is 8.65. The free energy difference between burying the uncharged and charged forms of the groups is modest, about 6 kcal/mol. We also found that epsilon(app) for I92K and I92E is in the range approximately 10-12, instead of 2-4 commonly used to represent the protein interior. Side-chains 92E and 92K were uncharged under the conditions of the X-ray experiment. Both are buried completely inside the well-defined hydrophobic core of the variant proteins without forming salt-bridges or hydrogen bonds to other functional groups of the proteins. Under cryo conditions 92E shows a chain of four water molecules, which hydrate one oxygen atom of the carboxyl group of the glutamic acid. Two other water molecules, which are present in the wild-type at all temperatures, are also connected to the water ring observed inside the hydrophobic core. The ready burial of water with an uncharged E92 raises the possibility that solvent excursions into the interior also take place in the wild-type protein, but in a random, dynamic way not detectable by crystallography. Such transient excursions could increase the average polarity, and thus epsilon(app), of the protein interior.  相似文献   

14.
The affinity label 5'-p-(fluorosulfonyl)benzoyl adenosine modifies rabbit muscle phosphofructokinase to the extent of one group/subunit. Modification appears to occur at a binding site specific for AMP, cyclic AMP, and ADP, i.e. those adenine nucleotides which are activators under conditions where regulatory kinetic behavior is obtained. The consequences of the modification are consistent with the model proposed previously for correlation between the pK of specific ionizable groups, regulatory kinetic behavior, ligand binding, and the reversible cold inactivation of the enzyme (Frieden, C., Gilbert. H. R., and Bock, P. E. (1976) J. Biol. Chem. 251, 5644-5647). Thus, the modification shifts the apparent pK of the essential ionizable groups from 6.9 to 6.4 at 25 degrees C, with the result that regulatory kinetic behavior at pH 6.9 and 25 degrees C is lost. Furthermore, the apparent affinity of a site (other than the active site) for ATP, as measured by ATP-dependent quenching of intrinsic protein fluorescence at pH 6.9 and 25 degrees C, is decreased by the modification. Regulatory kinetic behavior for both substrates is obtained with the modified enzyme at a lower pH, consistent with the downward shift in the pK of the ionizable groups, but sensitivity to cAMP activation is abolished by the modification. The loss of regulatory kinetic behavior upon modification of sulfhydryl groups does not appear to be the same as that due to modification by the affinity label.  相似文献   

15.
L-Pyroglutamate (PGA) is naturally occurring from L-glutamate solution with accelerated formation rate under high temperature and low pH. Even though PGA has been identified as a neurotoxic agent on brain cells, the effect of PGA on the growth of microorganisms is rarely known. Here various kinds of microorganisms differing in their optimal growth temperature, pH, phylogeny, and isolated biotope were investigated for the effect of PGA. We found that growth of thermoacidophiles, including both archaea and bacteria, was seriously inhibited by the presence of PGA, and the extent of the inhibitory effect was closely related with growth temperature and pH. Interestingly, only microbes that grow at high temperature and low pH are inhibited by PGA, while this compound may stimulate growth rates of organisms that live at neutral pH and low temperature.  相似文献   

16.
The active site of isomalto-dextranase from Arthrobacter globiformis was investigated by kinetic and chemical-modification methods. The ionization constants, pKe1 and pKe2, of the essential ionizable groups 1 and 2 of the free enzyme were 3.3 and 6.3 for dextran T2000 and 3.5 and 6.1 for isomaltotriose. The pKel and pKe2 both shifted to higher pH when the dielectric constant of the reaction mixture decreased. The heats of ionization for groups 1 and 2 were 0 kcal/mol or less with both substrates. These kinetic results suggested that the ionizable groups essential for the enzyme activity were carboxyl and carboxylate. Modification with 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide, modifying carboxyl residues specifically, resulted in inactivation of the enzyme, and isomaltotriose protected the enzyme against such inactivation. These findings also indicated that the carboxyl groups were essential to the enzyme activity.  相似文献   

17.
The membrane-traversing subunit c parallel from the F0 part of the ATP synthase molecule has been studied in chloroform/methanol by high-resolution 1H n.m.r. Various one-dimensional and two-dimensional techniques have been used for assignment purposes, some NOE connectivities were established and some 3JHN alpha coupling constants were measured from spin--echo experiments. The effects of varying pH, solvent composition, lanthanide concentration and temperature have been investigated. Evidence is presented that the molecule has extensive alpha-helical segments, and the hairpin structure suggested by other groups is supported by our n.m.r. data. Only one ionizable group, assigned to the C-terminal carboxyl, is observed to titrate in the pH range 2 to 10; so the conserved residue, Asp61, which binds dicyclohexylcarbodiimide, presumably has (at least in this solvent system) an abnormally high pK value.  相似文献   

18.
The diffuse wide angle x-ray scattering (WAXS) of polyglutamic acid (PGA) in solution was studied using an x-ray diffractometer with small aperture of the primary beam. The scattering curve was recorded at an angular interval from (article: see text). The experimental scattering intensity of PGA with alpha-helical CD spectrum showed a maximum at 14.4 nm-1. Unordered PGA in solution yielded no maximum at this scattering angle. The studies have proved that the scattering theory can be applied to globular proteins in solution as well as to chain molecules in solution in this angular interval. The differences between the calculated scattering curves and the experimental curves indicate minor movements of the side chains of PGA in solutions and slight structuring of the solvent at the surface of the polypeptide chain.  相似文献   

19.
A theory of site–site interaction due to proton/charge fluctuations in protein molecules has been developed. It is shown that, with the proper geometric configuration of identical ionizable groups on matching sites, a specific attraction may be established. This attractive force has a bell-shaped pH dependence and is maximal close to the pK of the groups involved. Various types of protein interactions are examined in the light of this theory.  相似文献   

20.
One clone (ACPGA001) exhibiting penicillin G acylase (PGA) activity was screened from a metagenomic library by using a medium containing penicillin G. A novel PGA gene from the inserted fragment of ACPGA001 was obtained by sequencing. The amino acid sequence of ACPGA001 PGA exhibited <33 % similarity to PGAs retrieved from GenBank. This gene was expressed in Escherichia coli M15 and the recombinant protein was purified and characterized. The ACPGA001 PGA exhibited a maximum activity at 60 °C and showed high activity at pH 4–10 with an optimum pH of 8.0. This enzyme was stable at 40 °C for 70 min with a half-life of 60 min at 55 °C. These beneficial characteristics of ACPGA001 PGA provide some advantages for the potential application of ACPGA001 PGA in industry.  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