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 46 毫秒
1.
Summary Commercially obtained cystine binding protein (CBP), an osmotic shock protein ofEscherichia coli, was studied in an effort to determine its binding characteristics. Sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS/PAGE) analysis of commercially obtained CBP showed three protein bands. N-terminal amino acid microsequencing and subsequent computer search revealed that the sequence of one of these proteins (25-kDa) was nearly identical to histidine binding protein (HisJ) ofSalmonella typhimurium. Purification of CBP by HPLC yielded four protein peaks, of which one bound histidine exclusively. Binding was maximal at pH 5.0 to 6.0, at 4°C, did not require calcium or magnesium ions and was not inhibited by reduction of CBP disulfide bonds. Amino acids other than histidine or cystine did not bind to CBP. These data show that commercially available CBP is not a homogenous protein; it contains a histidine as well as a cystine binding component.  相似文献   

2.
Subversion of the plasminogen activation system is implicated in the virulence of group A streptococci (GAS). GAS displays receptors for the human zymogen plasminogen on the cell surface, one of which is the plasminogen-binding group A streptococcal M-like protein (PAM). The plasminogen binding domain of PAM is highly variable, and this variation has been linked to host selective immune pressure. Site-directed mutagenesis of full-length PAM protein from an invasive GAS isolate was undertaken to assess the contribution of residues in the a1 and a2 repeat domains to plasminogen binding function. Mutagenesis to alanine of key plasminogen binding lysine residues in the a1 and a2 repeats (Lys98 and Lys111) did not abrogate plasminogen binding by PAM nor did additional mutagenesis of Arg101 and His102 and Glu104, which have previously been implicated in plasminogen binding. Plasminogen binding was only abolished with the additional mutagenesis of Arg114 and His115 to alanine. Furthermore, mutagenesis of both arginine (Arg101 and Arg114) and histidine (His102 and His115) residues abolished interaction with plasminogen despite the presence of Lys98 and Lys111 in the binding repeats. This study shows for the first time that residues Arg101, Arg114, His102, and His115 in both the a1 and a2 repeat domains of PAM can mediate high affinity plasminogen binding. These data suggest that highly conserved arginine and histidine residues may compensate for variation elsewhere in the a1 and a2 plasminogen binding repeats, and may explain the maintenance of high affinity plasminogen binding by naturally occurring variants of PAM.  相似文献   

3.
Escherichia coli HisJ is a type II periplasmic binding protein that functions to reversibly capture histidine and transfer it to its cognate inner membrane ABC permease. Here, we used NMR spectroscopy to determine the structure of apo-HisJ (26.5 kDa) in solution. HisJ is a bilobal protein in which domain 1 (D1) is made up of two noncontiguous subdomains, and domain 2 (D2) is expressed as the inner domain. To better understand the roles of D1 and D2, we have isolated and characterized each domain separately. Structurally, D1 closely resembles its homologous domain in apo- and holo-HisJ, whereas D2 is more similar to the holo-form. NMR relaxation experiments reveal that HisJ becomes more ordered upon ligand binding, whereas isolated D2 experiences a significant reduction in slower (millisecond to microsecond) motions compared with the homologous domain in apo-HisJ. NMR titrations reveal that D1 is able to bind histidine in a similar manner as full-length HisJ, albeit with lower affinity. Unexpectedly, isolated D1 and D2 do not interact with each other in the presence or absence of histidine, which indicates the importance of intact interdomain-connecting elements (i.e. hinge regions) for HisJ functioning. Our results shed light on the binding mechanism of type II periplasmic binding proteins where ligand is initially bound by D1, and D2 plays a supporting role in this dynamic process.  相似文献   

4.
The Escherichia coli histidine binding protein HisJ is a type II periplasmic binding protein (PBP) that preferentially binds histidine and interacts with its cytoplasmic membrane ABC transporter, HisQMP2, to initiate histidine transport. HisJ is a bilobal protein where the larger Domain 1 is connected to the smaller Domain 2 via two linking strands. Type II PBPs are thought to undergo “Venus flytrap” movements where the protein is able to reversibly transition from an open to a closed conformation. To explore the accessibility of the closed conformation to the apo state of the protein, we performed a set of all‐atom molecular dynamics simulations of HisJ starting from four different conformations: apo‐open, apo‐closed, apo‐semiopen, and holo‐closed. The simulations reveal that the closed conformation is less dynamic than the open one. HisJ experienced closing motions and explored semiopen conformations that reverted to closed forms resembling those found in the holo‐closed state. Essential dynamics analysis of the simulations identified domain closing/opening and twisting as main motions. The formation of specific inter‐hinge strand and interdomain polar interactions contributed to the adoption of the closed apo‐conformations although they are up to 2.5‐fold less prevalent compared with the holo‐closed simulations. The overall sampling of the closed form by apo‐HisJ provides a rationale for the binding of unliganded PBPs with their cytoplasmic membrane ABC transporters. Proteins 2014; 82:386–398. © 2013 Wiley Periodicals, Inc.  相似文献   

5.
Efficient release of ligands from the Ca(2+)-dependent carbohydrate-recognition domain (CRD) of the hepatic asialoglycoprotein receptor at endosomal pH requires a small set of conserved amino acids that includes a critical histidine residue. When these residues are incorporated at corresponding positions in an homologous galactose-binding derivative of serum mannose-binding protein, the pH dependence of ligand binding becomes more like that of the receptor. The modified CRD displays 40-fold preferential binding to N-acetylgalactosamine compared with galactose, making it a good functional mimic of the asialoglycoprotein receptor. In the crystal structure of the modified CRD bound to N-acetylgalactosamine, the histidine (His(202)) contacts the 2-acetamido methyl group and also participates in a network of interactions involving Asp(212), Arg(216), and Tyr(218) that positions a water molecule in a hydrogen bond with the sugar amide group. These interactions appear to produce the preference for N-acetylgalactosamine over galactose and are also likely to influence the pK(a) of His(202). Protonation of His(202) would disrupt its interaction with an asparagine that serves as a ligand for Ca(2+) and sugar. The structure of the modified CRD without sugar displays several different conformations that may represent structures of intermediates in the release of Ca(2+) and sugar ligands caused by protonation of His(202).  相似文献   

6.
Shigella dysentriae and other Gram‐negative human pathogens are able to use iron from heme bound to hemoglobin for growing. We solved at 2.6 Å resolution the 3D structure of the TonB‐dependent heme/hemoglobin outer membrane receptor ShuA from S. dysenteriae. ShuA binds to hemoglobin and transports heme across the outer membrane. The structure consists of a C‐terminal domain that folds into a 22‐stranded transmembrane β‐barrel, which is filled by the N‐terminal plug domain. One distal histidine ligand of heme is located at the apex of the plug, exposed to the solvent. His86 is situated 9.86 Å apart from His420, the second histidine involved in the heme binding. His420 is in the extracellular loop L7. The heme coordination by His86 and His420 involves conformational changes. The comparisons with the hemophore receptor HasR of Serratia marcescens bound to HasA‐Heme suggest an extracellular induced fit mechanism for the heme binding. The loop L7 contains hydrophobic residues which could interact with the hydrophobic porphyring ring of heme. The energy required for the transport by ShuA is derived from the proton motive force after interactions between the periplasmic N‐terminal TonB‐box of ShuA and the inner membrane protein, TonB. In ShuA, the TonB‐box is buried and cannot interact with TonB. The structural comparisons with HasR suggest its conformational change upon the heme binding for interacting with TonB. The signaling of the heme binding could involve a hydrogen bond network going from His86 to the TonB‐box. Proteins 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

7.
L A LeBrun  B V Plapp 《Biochemistry》1999,38(38):12387-12393
The rate of association of NAD(+) with wild-type horse liver alcohol dehydrogenase (ADH) is maximal at pH values between pK values of about 7 and 9, and the rate of NADH association is maximal at a pH below a pK of 9. The catalytic zinc-bound water, His-51 (which interacts with the 2'- and 3'-hydroxyl groups of the nicotinamide ribose of the coenzyme in the proton relay system), and Lys-228 (which interacts with the adenosine 3'-hydroxyl group and the pyrophosphate of the coenzyme) may be responsible for the observed pK values. In this study, the Lys228Arg, His51Gln, and Lys228Arg/His51Gln (to isolate the effect of the catalytic zinc-bound water) mutations were used to test the roles of the residues in coenzyme binding. The steady state kinetic constants at pH 8 for the His51Gln enzyme are similar to those for wild-type ADH. The Lys228Arg and Lys228Arg/His51Gln substitutions decrease the affinity for the coenzymes up to 16-fold, probably due to altered interactions with the arginine at position 228. As determined by transient kinetics, the rate constant for association of NAD(+) with the mutated enzymes no longer decreases at high pH. The pH profile for the Lys228Arg enzyme retains the pK value near 7. The His51Gln and Lys228Arg/His51Gln substitutions significantly decrease the rate constants for NAD(+) association, and the pH dependencies show that these enzymes bind NAD(+) most rapidly at a pH above pK values of 8. 0 and 9.0, respectively. It appears that the pK of 7 in the wild-type enzyme is shifted up by the H51Q substitutions, and the resulting pH dependence is due to the deprotonation of the catalytic zinc-bound water. Kinetic simulations suggest that isomerization of the enzyme-NAD(+) complex is substantially altered by the mutations. In contrast, the pH dependencies for NADH association with His51Gln, Lys228Arg, and Lys228Arg/His51Gln enzymes were the same as for wild-type ADH, suggesting that the binding of NAD(+) and the binding of NADH are controlled differently.  相似文献   

8.
Yamamoto T  Chen HC  Guigard E  Kay CM  Ryan RO 《Biochemistry》2008,47(44):11647-11652
The release of ligand from the low-density lipoprotein receptor (LDLR) has been postulated to involve a "histidine switch"-induced intramolecular rearrangement that discharges bound ligand. A recombinant soluble low-density lipoprotein receptor (sLDLR) was employed in ligand binding experiments with a fluorescently tagged variant apolipoprotein E N-terminal domain (apoE-NT). Binding was monitored as a function of fluorescence resonance energy transfer (FRET) from excited Trp residues in sLDLR to an extrinsic fluorophore covalently attached to Trp-null apoE3-NT. In binding experiments with wild-type (WT) sLDLR, FRET-dependent AEDANS fluorescence decreased as the pH was lowered. To investigate the role of His190, His562, and His586 in sLDLR in pH-dependent ligand binding and discharge, site-directed mutagenesis studies were performed. Compared to WT sLDLR, triple His --> Ala mutant sLDLR displayed attenuated pH-dependent ligand binding and a decreased level of ligand release as a function of low pH. When these His residues were substituted for Lys, the positively charged side chain of which does not ionize over this pH range, ligand binding was nearly abolished at all pH values. When sequential His to Lys mutants were examined, the evidence suggested that His562 and His586 function cooperatively. Whereas the sedimentation coefficient for WT sLDLR increased when the pH was reduced from 7 to 5, no such change occurred in the case of the triple Lys mutant receptor or a His562Lys/His586Lys double mutant receptor. The data support the existence of a cryptic, histidine side chain ionization-dependent alternative ligand that modulates ligand discharge via conformational reorganization.  相似文献   

9.
Site-directed mutagenesis has been utilized to examine the nature of the interaction of the histidine-binding protein (HisJ) with the membrane-bound components of the histidine transport system. In order to examine a region of the HisJ protein involved in the interaction with the membrane components, a number of charged amino acids in the vicinity of the genetically isolated interaction mutant hisJ5625 (R176C) were mutated. It was found that residues Asp171, Arg176, and Asp178 could be independently altered without affecting the histidine-binding affinity of the HisJ protein. However, the alteration of residues Asp171 and Arg176 greatly reduced the interaction of the HisJ protein with the membrane protein complex, whereas altering residue Asp178 had no effect on this interaction. Simultaneously, altering residues Asp183 and Glu184 resulted in a completely defective protein. The ability of a his-J5625 suppressor HisP protein (HisP(T205A)) to suppress the newly created site-directed mutants was also examined. This suppressor demonstrated specificity toward the amino acid present at position 176 and was also able the suppress the mutation created at position 171.  相似文献   

10.
Hainantoxin‐IV (HNTX‐IV), isolated from the venom of the spider Ornithoctonus hainana, is a specific antagonist of tetrodotoxin‐sensitive (TTX‐S) voltage‐gated sodium channels in rat dorsal root ganglion (DRG) cells. It adopts an inhibitor cystine knot motif, and structural analysis revealed a positively charged patch consisting of Arg26, Lys27, His28, Arg29 and Lys32 distributed on its molecular surface. Our previous study demonstrated that Lys27 and Arg29 but not Arg26 were critical residues for HNTX‐IV binding to TTX‐S sodium channels. In the present study, we examined the roles of His28 and Lys32 in the interaction of HNTX‐IV with its target. Two mutants, HNTX‐IV‐H28D and HNTX‐IV‐K32A, were generated by solid‐phase chemical synthesis and purified by reverse‐phase HPLC after refolding and oxidation, yielding two compounds of high purity with monoisotopic masses of 3962.66 and 3927.70 Da, respectively, as determined by MALDI‐TOF mass spectrometry. This indicated the presence of six cysteine residues forming three disulfide bonds. Moreover, circular dichroism spectroscopy analysis demonstrated that the substitution of His28 or Lys32 did not affect the overall structure of HNTX‐IV. The inhibitory activity of HNTX‐IV‐H28D and HNTX‐IV‐K32A against TTX‐S sodium channels in rat DRG cells was analyzed by whole‐cell patch‐clamp technique. The IC50 values for the mutants were 0.57 and 5.80 μM (17‐fold and 170‐fold lower than the activity of the native toxin), indicating that His28 and Lys32 may be important for the inhibitory activity of HNTX‐IV. Taken together, our results suggest that the positively charged patch might be the binding site for the interaction of HNTX‐IV with TTX‐S sodium channels. These findings might contribute to the elucidation of the structure and function relationship of HNTX‐IV. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
Bandarian V  Matthews RG 《Biochemistry》2001,40(16):5056-5064
Cobalamin-dependent methionine synthase (MetH) catalyzes the methylation of homocysteine using methyltetrahydrofolate as the methyl donor. The cobalamin cofactor serves as an intermediate carrier of the methyl group from methyltetrahydrofolate to homocysteine. In the two half-reactions that comprise turnover for MetH, the cobalamin is alternatively methylated by methyltetrahydrofolate and demethylated by homocysteine to form methionine. Upon binding to the protein, the usual dimethylbenzimidazole ligand is replaced by the imidazole side chain of His759 [Drennan, C. L., Huang, S., Drummond, J. T., Matthews, R. G., and Ludwig, M. L. (1994) Science 266, 1669-1674]. Despite the ligand replacement that accompanies binding of cobalamin to the holo-MetH protein, a MetH(2-649) fragment of methionine synthase that contains the regions that bind homocysteine and methyltetrahydrofolate utilizes exogenously supplied cobalamin in methyl transfer reactions akin to those of the catalytic cycle. However, the interactions of MetH(2-649) with endogenous cobalamin are first order in cobalamin, while the half-reactions catalyzed by the holoenzyme are zero order in cobalamin, so rate constants for reactions of bound and exogenous cobalamins cannot be compared. In this paper, we investigate the catalytic rate enhancements generated by binding cobalamin to MetH after dividing the protein in half and reacting MetH(2-649) with a second fragment, MetH(649-1227), that harbors the cobalamin cofactor. The second-order rate constant for demethylation of methylcobalamin by Hcy is elevated 60-fold and that for methylation of cob(I)alamin is elevated 120-fold. Thus, binding of cobalamin to MetH is essential for efficient catalysis.  相似文献   

12.
The effect of a continuous duodenal infusion of L‐histidine (His) (8 g/d) on the retention of nitrogen was investigated in two experiments (I, II), each of which was carried out using two young bulls. In Exps. I and II, the animals (150–250 kg BW) were fitted with a re‐entrant cannula in the proximal duodenum and were fed diets containing 125 g CP/kg DM and 11.5 MJ ME/kg DM. A third experiment (III) using two young bulls (140–200 kg BW) fitted with a simple T‐cannula was carried out infusing 6 g L‐His. The animals were fed a low protein diet (94 g CP/kg DM and 11 MJ ME/kg DM). The study was done to find out whether or not L‐His is the first limiting amino acid (AA) for growing ruminants.

N retention was 28 and 31, 38 and 38, 22 and 24 g/d without L‐His infusion and with L‐His infusion for Exps. I, II and III, respectively. Both in the experiments with a standard protein supply (I, II) and in the experiment with reduced protein supply (III), no significant differences were found between periods with and without infusion of L‐His. The utilisation of duodenal NAN varied between 39% and 50% and was also not significantly influenced by the duodenal infusion of L‐His. No significant effect was observed on the flow of AA into the duodenum. The faecal excretion of AA was also not significantly influenced by the infusion of L‐His. The utilisation of individual amino acids as calculated by the ratio of retained AA to intestinal apparently digested AA, did not differ significantly following the duodenal infusion of L‐His. As expected, the utilisation of His decreased. Of the different essential AA, L‐His was the most utilised (80%) followed by Arg (72%), Met (60%), Leu (45%) and Lys (44%), during periods without supplementation of L‐His.

It is concluded that the intestinal supply of L‐His from the basal diet was sufficient for the potential growth level of animals under these experimental conditions. In all AA present at the proximal duodenum, L‐His could have at first a limiting effect on the performance of growing young bulls with high body gain. Arg and Met, but not Lys, could be second or co‐limiting AA.  相似文献   

13.
We have previously provided compelling evidence that human recombinant interleukin 2 (IL-2) binds to the sulfated polysaccharides heparin, highly sulfated heparan sulfate and fucoidan. Here we show that IL-2 binding is dependent on heparin chain length, but with fragments as small as 15-mers retaining binding activity. The addition of exogenous heparin has no effect on the in vitro biological activity of IL-2. In addition soluble IL-2 receptor alpha and beta polypeptides do not compete with heparin for the binding of IL-2. IL-2 bound by heparin is still recognized by two IL-2 specific monoclonal antibodies, 3H9 and H2- 8, whose epitopes lie in the amino terminal region. Murine IL-2 unlike its human counterpart fails to bind to heparin. Human IL-2 analogs with single amino acid substitutions at positions Lys43, Thr51, and Gln126 analogs no longer bind to heparin. By contrast the Arg38Ala analog retains heparin full heparin binding activity. These experimental findings together with molecular modeling studies suggest two putative heparin binding sites on human IL-2, one involving four basic residues, Lys48, Lys49, Lys54, and His55, and the other being a discontinuous site comprising Lys43, Lys64, Arg81, and Arg83. Neither of these two clusters is completely conserved in murine IL-2. Overall our data suggest that the binding of human IL-2 to heparin and heparan sulfate does not interfere with IL-2/IL-2 receptor interactions. Therefore, binding to glycosaminoglycan may be a mechanism for retaining the cytokine in an active form close to its site of secretion in the tissue, thus favoring a paracrine role for IL-2.   相似文献   

14.
Hepatic-type fatty-acid-binding protein (hFABP) from the cytosol of bovine liver is a 14.4-kDa neutral protein with a blocked N-terminus and a disulfide system located on the surface of the protein. It binds two molecules of fatty acid in one binding site, apparent dissociation constants of the oleic acid/hFABP complex are 0.24 microM and 2.15 microM. Computer analysis of circular dichroic spectra predicts that hFABP contains about 12% alpha-helix, 45% beta-structure, 15% beta-turn and 27% unordered structure. Ellipticities indicative of secondary structure are not affected by fatty acid binding. Cationic amino acid residues of hFABP (1 His, 15 Lys, 2 Arg) were screened for ionic fatty acid/protein interactions. His was excluded, as 1H-NMR analysis of His-C2 and His-C4 protons indicated that binding of oleic acid shifts the pK of His from 6.9 to 7.1 only in hFABP with the disulfide system in the oxidized state; acylation of His with diethylpyrocarbonate does not affect the binding of the fatty acid. Acetylation of Lys reduces binding marginally, whereas modification of Arg with phenylglyoxal lowers the binding activity by 65%. From 1H-NMR investigations, conformational changes within the protein, due to a sort of disaggregation of hFABP upon fatty acid binding, were derived. Most of the proton resonances sharpen up with ligand binding, and some of the methyl resonances shift positions, possibly because they are directly involved in the fatty acid/protein interaction.  相似文献   

15.
We have examined the effects of active site residues on ligand binding to the heme iron of mouse neuroglobin using steady-state and time-resolved visible spectroscopy. Absorption spectra of the native protein, mutants H64L and K67L and double mutant H64L/K67L were recorded for the ferric and ferrous states over a wide pH range (pH 4-11), which allowed us to identify a number of different species with different ligands at the sixth coordination, to characterize their spectroscopic properties, and to determine the pK values of active site residues. In flash photolysis experiments on CO-ligated samples, reaction intermediates and the competition of ligands for the sixth coordination were studied. These data provide insights into structural changes in the active site and the role of the key residues His64 and Lys67. His64 interferes with exogenous ligand access to the heme iron. Lys67 sequesters the distal pocket from the solvent. The heme iron is very reactive, as inferred from the fast ligand binding kinetics and the ability to bind water or hydroxyl ligands to the ferrous heme. Fast bond formation favors geminate rebinding; yet the large fraction of bimolecular rebinding observed in the kinetics implies that ligand escape from the distal pocket is highly efficient. Even slight pH variations cause pronounced changes in the association rate of exogenous ligands near physiological pH, which may be important in functional processes.  相似文献   

16.
HisJ is a histidine binding subunit of the histidine permease, which exists in the outer membrane of Gram-negative bacteria. In order to incorporate the periplasmic histidine into the cell, HisJ captures histidine in the periplasm, and transfers the histidine to the transmembrane complex of histidine permease that is an ABC transporter. We established the backbone resonance assignments of 1H/13C/15N-labeled HisJ from Escherichia coli, in the histidine-bound and unbound states.  相似文献   

17.
L-Tryptophan is the least abundant essential amino acid in humans. Indoleamine 2,3-dioxgyenase (IDO) is a cytosolic heme protein which, together with the hepatic enzyme tryptophan 2,3-dioxygenase, catalyzes the first and rate-limiting step in the major pathway of tryptophan metabolism, the kynurenine pathway. The physiological role of IDO is not fully understood but is of great interest, because IDO is widely distributed in human tissues, can be up-regulated via cytokines such as interferon-gamma, and can thereby modulate the levels of tryptophan, which is vital for cell growth. To identify which amino acid residues are important in substrate or heme binding in IDO, site-directed mutagenesis of conserved residues in the IDO gene was undertaken. Because it had been proposed that a histidine residue might be the proximal heme ligand in IDO, mutation to alanine of the three highly conserved histidines His16, His303, and His346 was conducted. Of these, only His346 was shown to be essential for heme binding, indicating that this histidine residue may be the proximal ligand and suggesting that neither His303 nor His16 act as the proximal ligand. Site-directed mutagenesis of Asp274 also compromised the ability of IDO to bind heme. This observation indicates that Asp274 may coordinate to heme directly as the distal ligand or is essential in maintaining the conformation of the heme pocket.  相似文献   

18.
Glycerol-3-phosphate 1-acyltransferase is a soluble chloroplast enzyme involved in glycerol-lipid biosynthesis associated with chilling resistance in plants (). Resistance is associated with higher selectivity for unsaturated acyl substrates over saturated ones. In vitro substrate selectivity assays performed under physiologically relevant conditions have been established that discriminate between selective and non-selective forms of the enzyme. A mutation, L261F, in the squash protein converts it from a non-selective enzyme into a selective one. The mutation lies within 10 A of the predicted acyl binding site and results in a higher K(m) for 16:0 acyl carrier protein (ACP). Site-directed mutagenesis was used to determine the importance of four residues, Arg(235), Arg(237), Lys(193), and His(194), implicated to be involved in binding of the phosphate group of glycerol 3-phosphate to the enzyme. All the proteins were highly homologous in structure to the wild type enzyme. Mutations in Arg(235), Arg(237), and Lys(193) resulted in inactive enzyme, while His(194) had reduced catalytic activity. The mutant proteins retained the ability to bind stoichiometric quantities of acyl-ACPs supporting the potential role of these residues in glycerol 3-phosphate binding.  相似文献   

19.
Summary. The interaction of free amino acids with the corn protein zein was studied by thin-layer chromatography carried out on cellulose layers covered with zein and the effect of pH and salts on the strength of interaction was elucidated. Only the binding of Arg, His, Lys, Orn and Trp to zein was verified, other amino acids were not retained. Retention of Arg, His, Lys and Orn decreased linearly with increasing concentration of salts the mobile phase indicating the hydrophilic character of amino acid–zein interaction. Both alkaline and acidic pH influenced the strength of binding. Principal component analysis indicated the different character of the influence of pH and salts on the interaction. The results suggest that these amino acid residues may account for the binding of other peptides and proteins to zein.  相似文献   

20.
The influence of three alkaline amino acids, l ‐lysine (L ‐Lys), l ‐arginine (L ‐Arg) and l ‐histidine (L‐His), on the aggregation behaviour of sodium deoxycholate (NaDC) in phosphate buffer, pH 7.0, was studied at 25 °C. The fluorescence probe technique based on pyrene was employed to determine accurately the critical aggregation concentration (cac), polarity of the microenvironment and aggregation numbers for the NaDC aggregates. The added amino acids can effectively reduce the cac values and micropolarity of NaDC, indicating that it is easier for NaDC to aggregate in a compact manner in the presence of amino acids. The aggregation numbers of NaDC were increased, indicating that more NaDC molecules connect together to form stable aggregates. The performance of L‐Arg is similar to that of L‐His, and both have a smaller effect on the above parameters than L ‐Lys. In view of this, it may be inferred that both electrostatic and hydrophobic interaction are responsible for the interaction between NaDC and amino acids in aqueous solution. Copyright © 2011 John Wiley & Sons, Ltd.  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