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.
The relationship between the two principal ligand binding sites, sites I and II, on human serum albumin (HSA) was quantitatively and qualitatively examined by equilibrium dialysis and fluorescence spectroscopy. Among the three subsite markers to site I, only the binding of dansyl-L-asparagine (DNSA), which is a subsite Ib marker (K. Yamasaki et al., Biochim. Biophys. Acta 1295 (1996) 147), was inhibited by the simultaneous binding of a site II ligand, such as ibuprofen and diazepam. This indicates that, in contrast to subsite Ib, subsites Ia and Ic do not strongly interact with site II. The thermodynamic characteristics for the coupling reaction between DNSA and ibuprofen and between DNSA and diazepam, which gave positive coupling free energies and negative values for both coupling enthalpy and entropy, indicated that the reaction process was entropically driven. Increase of pH from 6.5 to 8.2 caused an increase in coupling constant and entropy for the mutual antagonism between DNSA and the site II ligands on binding to HSA. The site II ligand-induced red-shift of lambda(max) and solvent accessibility of DNSA in subsite Ib were decreased when the albumin molecule was isomerized from the neutral (N) to the base (B) conformation in the physiological pH region. Based on these findings, we conclude that a 'competitive' like strong allosteric regulation exists for the binding of these two ligands to the N conformer, whereas for the B conformer this interaction can be classified as nearly 'independent'. Since the distance between Trp-214, which resides within the site I subdomain, and Tyr-411, which is involved in site II, is increased by 6 A during the N-B transition (N.G. Hagag et al., Fed. Proc. 41 (1982) 1189), we propose a mechanism for the pH-dependent antagonistic binding between subsite Ib and site II, which involves the transmission of ligand-induced allosteric effects from one site to another site, modified by changes in the spatial relationship of sites I and II caused by the N-B transition.  相似文献   

2.
This study represents the first characterisation of the substrate-binding site of Bacillus licheniformis alpha-amylase (BLA). It describes the first subsite map, namely, number of subsites, apparent subsite energies and the dual product specificity of BLA. The product pattern and cleavage frequencies were determined by high-performance liquid chromatography, utilising a homologous series of chromophore-substituted maltooligosaccharides of degree of polymerisation 4-10 as model substrates. The binding region of BLA is composed of five glycone, three aglycone-binding sites and a 'barrier' subsite. Comparison of the binding energies of subsites, which were calculated with a computer program, shows that BLA has similarity to the closely related Bacillus amyloliquefaciens alpha-amylase.  相似文献   

3.
Arthrobotrys amerospora ATCC 34468 produced glucoamylase in a medium containing maize starch as carbon source. On native PAGE, crude glucoamylase showed three isoenzymes which were designated as Glu I, Glu II, Glu III according to their electrophoretic mobility. These were purified by column chromatography techniques. The energy of binding for each glucoamylase was calculated using Hiromi's kinetic based calculation. At subsite 1, the binding energies for Glu I, II and III were found to be negative.  相似文献   

4.
The role of binding subsite A, located at the terminal of the six binding subsites of hen egg-white lysozyme, in substrate binding and catalytic reactions was investigated by kinetic studies using a chemical modification method. Computer simulation showed that, although subsite A participates in the binding of the substrate, a decrease in the affinity of subsite A to the sugar residue does not cause a lowering of the rate of substrate consumption but changes the mode of the reaction by changing the distribution of the products formed. The binding free energies of subsites for Asp-101-modified lysozymes were estimated by data-fitting from the experimental time-courses. The contribution of Asp-101 in hen egg-white lysozyme to the substrate binding at subsite A was estimated to correspond to a binding free energy of about -3 kJ/mol, 30% of the total binding free energy of subsite A. Modification of Asp-101 affected not only the binding free energy of subsite A but also that of subsite C.  相似文献   

5.
This study characterizes the substrate-binding sites of human salivary alpha-amylase (HSA) and its Y151M mutant. It describes the first subsite maps, namely, the number of subsites, the position of cleavage sites and apparent subsite energies. The product pattern and cleavage frequencies were determined by high-performance liquid chromatography, utilizing a homologous series of chromophore-substituted maltooligosaccharides of degree of polymerization 3-10 as model substrates. The binding region of HSA is composed of four glycone and three aglycone-binding sites, while that of Tyr151Met is composed of four glycone and two aglycone-binding sites. The subsite maps show that Y151M has strikingly decreased binding energy at subsite (+2), where the mutation has occurred (-2.6 kJ/mol), compared to the binding energy at subsite (+2) of HSA (-12.0 kJ/mol).  相似文献   

6.
In the three-dimensional structure of a rice class I chitinase (OsChia1b) determined recently, a loop structure (loop II) is located at the end of the substrate-binding cleft, and is thus suggested to be involved in substrate binding. In order to test this assumption, deletion of the loop II region from the catalytic domain of OsChia1b and replacement of Trp159 in loop II with Ala were carried out. The loop II deletion and the W159A mutation increased hydrolytic activity not only towards (GlcNAc)6 but also towards polysaccharide substrates. Similar results were obtained for kcat/Km values determined for substrate reduced-(GlcNAc)5. The two mutations shifted the splitting positions in (GlcNAc)6 to the reducing end side, but the shift was less intensive in the Trp mutant. Theoretical analysis of the reaction time course indicated that sugar residue affinity at the +3 subsite was reduced from -2 kcal/mol to +0.5 kcal/mol by loop II deletion. Reduced affinity at the +3 subsite might enhance the release of product fragments, resulting in higher turnover and higher enzymatic activities. Thus, we concluded that loop II is involved in sugar residue binding at the +3 subsite, but that Trp159 itself appears to contribute only partly to sugar residue interaction at the subsite.  相似文献   

7.
The structure of the complex between a catalytically compromised family 10 xylanase and a xylopentaose substrate has been determined by X-ray crystallography and refined to 3.2 A resolution. The substrate binds at the C-terminal end of the eightfold betaalpha-barrel of Pseudomonas fluorescens subsp. cellulosa xylanase A and occupies substrate binding subsites -1 to +4. Crystal contacts are shown to prevent the expected mode of binding from subsite -2 to +3, because of steric hindrance to subsite -2. The loss of accessible surface at individual subsites on binding of xylopentaose parallels well previously reported experimental measurements of individual subsites binding energies, decreasing going from subsite +2 to +4. Nine conserved residues contribute to subsite -1, including three tryptophan residues forming an aromatic cage around the xylosyl residue at this subsite. One of these, Trp 313, is the single residue contributing most lost accessible surface to subsite -1, and goes from a highly mobile to a well-defined conformation on binding of the substrate. A comparison of xylanase A with C. fimi CEX around the +1 subsite suggests that a flatter and less polar surface is responsible for the better catalytic properties of CEX on aryl substrates. The view of catalysis that emerges from combining this with previously published work is the following: (1) xylan is recognized and bound by the xylanase as a left-handed threefold helix; (2) the xylosyl residue at subsite -1 is distorted and pulled down toward the catalytic residues, and the glycosidic bond is strained and broken to form the enzyme-substrate covalent intermediate; (3) the intermediate is attacked by an activated water molecule, following the classic retaining glycosyl hydrolase mechanism.  相似文献   

8.
A computer program has been evaluated for subsite map calculations of depolymerases. The program runs in windows and uses the experimentally determined bond cleavage frequencies (BCFs) for determination of the number of subsites, the position of the catalytic site and for calculation of subsite binding energies. The apparent free energy values were optimized by minimization of the differences of the measured and calculated BCF data. The program called suma (SUbsite Mapping of alpha-Amylases) is freely available for research and educational purposes via the Internet (E-mail: gyemant@tigris.klte.hu). The advantages of this program are demonstrated through alpha-amylases of different origin, e.g. porcine pancreatic alpha-amylase (PPA) studied in our laboratory, in addition to barley and rice alpha-amylases published in the literature. Results confirm the popular 'five subsite model' for PPA with three glycone and two aglycone binding sites. Calculations for barley alpha-amylase justify the '6 + 2 + (1) model' prediction. The binding area of barley alpha-amylase is composed of six glycone, two aglycone binding sites followed by a barrier subsite at the reducing end of the binding site. Calculations for rice alpha-amylase represent an entirely new map with a '(1) + 2 + 5 model', where '(1)' is a barrier subsite at the nonreducing end of the binding site and there are two glycone and five aglycone binding sites. The rice model may be reminiscent of the action of the bacterial maltogenic amylase, that is, suggesting an exo-mechanism for this enzyme.  相似文献   

9.
Asp176, Glu179 and Glu180 of Aspergillus awamori glucoamylase appeared by differential labeling to be in the active site. To test their functions, they were replaced by mutagenesis with Asn, Gln and Gln respectively, and kinetic parameters and pH dependencies of all enzyme forms were determined. Glu179----Gln glucoamylase was not active on maltose or isomaltose, while the kcat for maltoheptaose hydrolysis decreased almost 2000-fold and the KM was essentially unchanged from wild-type glucoamylase. The The Glu180----Gln mutation drastically increased the KM and moderately decreased the kcat with maltose and maltoheptaose, but affected isomaltose hydrolysis less. Difference in substrate activation energies between Glu180----Gln and wild-type glucoamylases indicate that Glu180 binds D-glucosyl residues in subsite 2. The Asp176----Asn substitution gave moderate increases and decreases in KM and kcat respectively, and therefore similar increases in activation energies for the three substrates. This and the differences in subsite binding energies between Asp176----Asn and wild-type glucoamylases suggest that Asp176 is near subsite 1, where it stabilizes the transition state and interacts with Trp120 at subsite 4. Glu179 and Asp176 are thus proposed as the general catalytic acid and base of pKa 5.9 and 2.7 respectively. The charged Glu180 contributes to the high pKa value of Glu179.  相似文献   

10.
To investigate the functional role of subsites E and F in lysozyme catalysis, Asn37 of hen egg-white lysozyme (HEL), which is postulated to participate in sugar residue binding at the right-sided subsite F through hydrogen bonding, was replaced by Ser or Gly by site-directed mutagenesis. The mutations of Asn37 neither significantly affected the binding constant for chitotriose nor the enzymatic activity toward the substrate glycol chitin. However, kinetic analysis with the substrate N-acetylglucosamine pentamer, (GlcNAc)(5), revealed that the conversion of Asn37 to Gly decreased the binding free energies for subsites E and F, while the conversion to Ser increased the substrate affinity at subsite F. It was further found that the rate constant of transglycosylation was reduced by these mutations. These results suggest that Asn37 is involved not only in substrate binding at subsite F but also in transglycosylation activity. No remarkable change in the tertiary structure except the side chain of the 37th residue was detected on X-ray analysis of the mutant proteins, indicating that the alterations in the enzymatic function between the wild type and mutant enzymes depend on limited structural change around the substitution site. It is thus speculated that the slight conformational difference in the side chain of position 37 may affect the substrate and acceptor binding at subsites E and F, leading to lower the efficiency of the transglycosylation activities of the mutant proteins.  相似文献   

11.
Subsite affinity maps of long substrate binding clefts in barley alpha-amylases, obtained using a series of maltooligosaccharides of degree of polymerization of 3-12, revealed unfavorable binding energies at the internal subsites -3 and -5 and at subsites -8 and +3/+4 defining these subsites as binding barriers. Barley alpha-amylase 1 mutants Y105A and T212Y at subsite -6 and +4 resulted in release or anchoring of bound substrate, thus modifying the affinities of other high-affinity subsites (-2 and +2) and barriers. The double mutant Y105A-T212Y displayed a hybrid subsite affinity profile, converting barriers to binding areas. These findings highlight the dynamic binding energy distribution and the versatility of long maltooligosaccharide derivatives in mapping extended binding clefts in alpha-amylases.  相似文献   

12.
Family II pyrophosphatases (PPases), recently found in bacteria and archaebacteria, are Mn(2+)-containing metalloenzymes with two metal-binding subsites (M1 and M2) in the active site. These PPases can use a number of other divalent metal ions as the cofactor but are inactive with Zn(2+), which is known to be a good cofactor for family I PPases. We report here that the Mg(2+)-bound form of the family II PPase from Streptococcus gordonii is nearly instantly activated by incubation with equimolar Zn(2+), but the activity thereafter decays on a time scale of minutes. The activation of the Mn(2+)-form by Zn(2+) was slower but persisted for hours, whereas activation was not observed with the Ca(2+)- and apo-forms. The bound Zn(2+) could be removed from PPase by prolonged EDTA treatment, with a complete recovery of activity. On the basis of the effect of Zn(2+) on PPase dimerization, the Zn(2+) binding constant appeared to be as low as 10(-12) M for S. gordonii PPase. Similar effects of Zn(2+) and EDTA were observed with the Mg(2+)- and apo-forms of Streptococcus mutans and Bacillus subtilis PPases. The effects of Zn(2+) on the apo- and Mg(2+)-forms of HQ97 and DE15 B. subtilis PPase variants (modified M2 subsite) but not of HQ9 variant (modified M1 subsite) were similar to that for the Mn(2+)-form of wild-type PPase. These findings can be explained by assuming that (a) the PPase tightly binds Mg(2+) and Mn(2+) at the M2 subsite; (b) the activation of the corresponding holoenzymes by Zn(2+) results from its binding to the M1 subsite; and (c) the subsequent inactivation of Mg(2+)-PPase results from Zn(2+) migration to the M2 subsite. The inability of Zn(2+) to activate apo-PPase suggests that Zn(2+) binds more tightly to M2 than to M1, allowing direct binding to M2. Zn(2+) is thus an efficient cofactor at subsite M1 but not at subsite M2.  相似文献   

13.
Phosphorus-31 nuclear magnetic resonance (NMR) studies on the two phosphorus nuclei of the phosphonium analogue (Me3P+CH2CH2OPO3(2-)) of phosphocholine are used to monitor the charged subsites in the phosphocholine-binding immunoglobulin A mouse myeloma M603. Comparison of the 270-MHz 1H NMR difference spectrum on addition of either this analogue or phosphocholine to M603 and the almost identical changes in the pKa values of the phosphate groups on binding to M603 confirm that the analogue is a good model for phosphocholine. The pKa of the phosphate groups is decreased by 0.5 unit on binding to M603, which is consistent with the phosphate group being hydrogen bonding to Tyr-33H and Arg-95L, as suggested from the X-ray structure, and also implies that the binding energies for the mono- and dianion are similar. The P+Me3 moiety is used to probe the electrostatic interactions in the choline subsite. Titration of the chemical shift of the phosphonium phosphorus reflects a group on the protein that has a pKa value of less than or equal to 5, which from the refined X-ray structure (D.R. Davies, personal communication) of the site is assigned to Asp-97L. The choline subsite is monitored by using 1H NMR difference spectra, which indicates that the subsite is highly aromatic as expected from the crystal structure that places Trp-107H and Tyr-100L in this subsite. The ring current interactions from these rings can account for the 1H NMR chemical shift data on choline.  相似文献   

14.
Reaction patterns for the hydrolysis of chromophoric glycosides from cello-oligosaccharides and lactose by the cellobiohydrolases (CBH I and CBH II) purified from Trichoderma reesei and Penicillium pinophilum were determined. They coincide with those found for the parent unsubstituted sugars. CBH I enzyme from both organisms attacks these substrates in a random manner. Turnover numbers are, however, low and do not increase appreciably as a function of the degree of polymerization of the substrates. The active-site topology of the CBH I from T. reesei was further probed by equilibrium binding experiments with cellobiose, cellotriose, lactose and some of their derivatives. These point to a single interaction site (ABC), spatially restricted as deduced from the apparent independency of the thermodynamic parameters. It appears that the putative subsite A can accommodate a galactopyranosyl or glucopyranosyl group, and subsite B a glucopyranosyl group, whereas in subsite C either a glucopyranosyl or a chromophoric group can be bound, scission occurring between subsites B and C. The apparent kinetic parameters (turnover numbers) for the hydrolysis of cello-oligosaccharides (and their derivatives) by the CBH II type enzyme increase as a function of chain length, indicative of an extended binding site (A-F). Its architecture allows for specific binding of beta-(1----4)-glucopyranosyl groups in subsites A, B and C. Binding of a chromophore in subsite C produces a non-hydrolysable complex. The thermodynamic interaction parameters of some ligands common to both type of enzyme were compared: these substantiate the conclusions reached above.  相似文献   

15.
Subsite mapping of enzymes. Depolymerase computer modelling.   总被引:2,自引:1,他引:1       下载免费PDF全文
We have developed a depolymerase computer model that uses a minimization routine. The model is designed so that, given experimental bond-cleavage frequencies for oligomeric substrates and experimental Michaelis parameters as a function of substrate chain length, the optimum subsite map is generated. The minimized sum of the weighted-squared residuals of the experimental and calculated data is used as a criterion of the goodness-of-fit for the optimized subsite map. The application of the minimization procedure to subsite mapping is explored through the use of simulated data. A procedure is developed whereby the minimization model can be used to determine the number of subsites in the enzymic binding region and to locate the position of the catalytic amino acids among these subsites. The degree of propagation of experimental variance into the subsite-binding energies is estimated. The question of whether hydrolytic rate coefficients are constant or a function of the number of filled subsites is examined.  相似文献   

16.
To elucidate how temperature effects subsite mapping of a thermostable alpha-amylase from Bacillus licheniformis (BLA), a comparative study was performed by using 2-chloro-4-nitrophenyl (CNP) beta-maltooligosides with degree of polymerisation (DP) 4-10 as model substrates. Action patterns, cleavage frequencies and subsite binding energies were determined at 50 degrees C, 80 degrees C and 100 degrees C. Subsite map at 80 degrees C indicates more favourable bindings compared to the hydrolysis at 50 degrees C. Hydrolysis at 100 degrees C resulted in a clear shift in the product pattern and suggests significant differences in the active site architecture. Two preferred cleavage modes were seen for all substrates in which subsite (+2) and (+3) were dominant, but CNP-G1 was never formed. In the preferred binding mode of shorter oligomers, CNP-G2 serves as the leaving group (79%, 50%, 59% and 62% from CNP-G4, CNP-G5, CNP-G6 and CNP-G7, respectively), while CNP-G3 is the dominant hydrolysis product from CNP-G8, CNP-G9, and CNP-Gl0 (62%, 68% and 64%, respectively). The high binding energy value (-17.5 kJ/mol) found at subsite (+2) is consistent with the significant formation of CNP-G2. Subsite mapping at 80 degrees C and 100 degrees C confirms that there are no further binding sites despite the presence of longer products.  相似文献   

17.
The association constants for the binding of various saccharides to hen egg-white lysozyme and human lysozyme have been measured by fluorescence titration. Among these are the oligosaccharides GlcNAc-beta(1 leads to 4)-MurNAc-beta(1 leads to 4)-GlcNAc-beta(1 leads to 4)-GlcNAc, GlcNAc-beta(1 leads to 4)-MurNAc-beta(1 leads to 4)-GlcNAc-beta(1 leads to 4)-N-acetyl-D-xylosamine, and GlcNAc-beta(1 leads to 4-GlcNAc-beta(1 leads to 4)-MurNAc, prepared here for the first time. The binding constants for saccharides which must have N-acetylmuramic acid, N-acetyl-D-glucosamine, or N-acetyl-D-xylosamine bound in subsite D indicate that there is no strain involved in the binding of N-acetyl-D-glycosamine in this site, and that the lactyl group of N-acetylmuramic acid (rather than the hydroxymethyl group) is responsible for the apparent strain previously reported for binding at this subsite. For hen egg-white lysozyme, the dependence of saccharide binding on pH or on a saturating concentration of Gd(III) suggests that the conformation of several of the complexes are different from one another and from that proposed for a productive complex. This is supported by fluorescence difference spectra of the various hen egg-white lysozyme-saccharide complexes. Human lysozyme binds most saccharides studied more weakly than the hen egg-white enzyme, but binds GlcNAc-beta(1 leads to 4)-MurNAc-beta(1leads to 4)-GlcNAc-beta(1 leads to 4)-MurNAc more strongly. It is suggested that subsite C of the human enzyme is "looser" than the equivalent site in the hen egg enzyme, so that the rearrangement of a saccharide in this subsite in response to introduction of an N-acetylmuramic acid residue into subsite D destabilizes the saccharide complexes of human lysozyme less than it does the corresponding hen egg-white lysozyme complexes. This difference and the differences in the fluorescence difference spectra of hen egg-white lysozyme and human lysozyme are ascribed mainly to the replacement of Trp-62 in hen egg-white lysozyme by Tyr-63 in the human enzyme. The implications of our findings for the assumption of superposition and additivity of energies of binding in individual subsites, and for the estimation of the role of strain in lysozyme catalysis, are discussed.  相似文献   

18.
Subsite mapping is a crucial procedure in the characterization of α-amylases (EC 3.2.1.1), which are extensively used in starch-based industries and in diagnosis of pancreatic and salivary glands disorders. A computer-aided method has been developed for subsite mapping of α-amylases, which substitutes the difficult, expensive, and time-consuming experimental determination of action patterns to crystal structures based energy calculations. Interaction energies between enzymes and carbohydrate substrates were calculated after short energy minimization by a molecular mechanics program. A training set of wild type and mutant amylases with known experimental action patterns of 13 enzymes of wide range of origin was used to set up the procedure. Calculations for training set resulted in good correlation in case of subsite binding energies (r2 = 0.827–0.929) and bond cleavage frequencies (r2 = 0.727–0.835). A set of eight novel barley amylase 1 mutants was used to test our model. Subsite binding energies were predicted with r2 = 0.502 correlation coefficient, while bond cleavage frequency prediction resulted in r2 = 0.538. Our computer-aided procedure may supplement the experimental subsite mapping methods to predict and understand characteristic features of α-amylases.  相似文献   

19.
Forty-one aromatic and aliphatic analogs of alpha-ketoglutarate were studied kinetically for their interaction with the alpha-ketoglutarate binding site of gamma-butyrobetaine hydroxylase obtained from Pseudomonas sp. AK1. Together, the compounds represent structural permutations probing the contribution of: 1) the C5 carboxyl group of alpha-ketoglutarate (domain I); 2) the C1-C2 keto acid moiety of alpha-ketoglutarate (domain II); 3) the distance between domains I and II; and 4) the spatial relationship of the two domains required for optimal interaction with the cosubstrate binding site. All compounds were competitive inhibitors for alpha-ketoglutarate (Km 0.018 mM). Functionally, two subsites of the cosubstrate binding site were evident: subsite I for polar interaction with the C5 carboxyl group, and subsite II, comprising of two distinct cis-oriented coordination sites of the catalytic ferrous ion which interact with the C1-C2 keto acid moiety. The most efficient inhibitors were pyridine 2,4-dicarboxylate (Ki 0.0002 mM) and 3,4-dihydroxybenzoate (Ki 0.0006 mM). Both compounds contain a carboxyl group and a chelating moiety corresponding to domains I and II of alpha-ketoglutarate, respectively. The fixed orientation of these groups in both analogs was used to assess intersubsite distance and spatial relationship required for optimal interaction with the cosubstrate binding site. Binding at subsite I and chelation at subsite II were indispensible for effective competitive inhibition. The distance between these two domains also helped determine whether attachment at the cosubstrate binding site would be catalytically productive. This was emphasized by the failure of either oxaloacetate or alpha-ketoadipinate to promote hydroxylation. Optimal interdomain distance, however, was not sufficient for cosubstrate utilization, as pyridine 2,4-dicarboxylate, with an interdomain distance identical to alpha-ketoglutarate in its staggered conformation, did not sustain hydroxylation. In the overall, these studies suggest that alpha-ketoglutarate utilization occurs in a ligand reaction at the active site ferrous ion of gamma-butyrobetaine hydroxylase. This is of particular interest since the delineated stereochemical mode of oxidative decarboxylation could generate the reactive oxo-iron species that was shown experimentally to promote gamma-butyrobetaine hydroxylation by an abstraction-recombination mechanism (Blanchard, J. S., and Englard, S. (1983) Biochemistry 22, 5922-5928; Englard, S., Blanchard, J. S., and Midelfort, C. F. (1985) Biochemistry 24, 1110-1116).  相似文献   

20.
Antibody IgE Lb4 interacts favorably with a large number of different compounds. To improve the current understanding of the structural basis of this vast cross-reactivity, the binding of three dinitrophenyl (DNP) amino acids (DNP-alanine, DNP-glycine, and DNP-serine) is investigated in detail by means of docking and molecular dynamics free energy simulations. Experimental binding energies obtained by isothermal titration microcalorimetry are used to judge the results of the computational studies. For all three ligands, the docking procedure proposes two plausible subsites within the binding region formed by the antibody CDR loops. By subsequent molecular dynamics simulations and calculations of relative free energies of binding, one of these subsites, a tyrosine-surrounded pocket, is revealed as the preferred point of complexation. For this subsite, results consistent with experimental observations are obtained; DNP-glycine is found to bind better than DNP-serine, and this, in turn, is found to bind better than DNP-alanine. The suggested binding mode makes it possible to explain both the moderate binding affinity and the differences in binding energy among the three ligands.  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