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.
ShK toxin is a structurally defined, 35-residue polypeptide which blocks the voltage-gated Kv1.3 potassium channel in T-lymphocytes and has been identified as a possible immunosuppressant. Our interest lies in the rational design and synthesis of type-III mimetics of protein and polypeptide structure and function. ShK toxin is a challenging target for mimetic design as its binding epitope consists of relatively weakly binding residues, some of which are discontinuous. We discuss here our investigations into the design and synthesis of 1st generation, small molecule mimetics of ShK toxin and highlight any principles relevant to the generic design of type-III mimetics of continuous and discontinuous binding epitopes. We complement our approach with attempted pharmacophore-based database mining.  相似文献   

2.
Animal venoms represent a rich source of pharmacologically active peptides that interact with ion channels. However, a challenge to discovering drugs remains because of the slow pace at which venom peptides are discovered and refined. An efficient autocrine‐based high‐throughput selection system was developed to discover and refine venom peptides that target ion channels. The utility of this system was demonstrated by the discovery of novel Kv1.3 channel blockers from a natural venom peptide library that was formatted for autocrine‐based selection. We also engineered a Kv1.3 blocker peptide (ShK) derived from sea anemone to generate a subtype‐selective Kv1.3 blocker with a long half‐life in vivo.  相似文献   

3.
The molecular mechanics Poisson-Boltzmann surface area (MM-PBSA) method combined with alanine-scanning mutagenesis is a very important tool for rational drug design. In this study, molecular dynamics (MD) and MM-PBSA were applied to calculate the binding free energy between the rat intestinal fatty acid binding protein (IFABP) and palmitic acid (PA) to gain insight to the interaction details. Equally spaced snapshots along the trajectory were chosen to perform the binding free energy calculation, which yields a result highly consistent with experimental value with a deviation of 0.4 kcal/mol. Computational alanine scanning was performed on the same set of snapshots by mutating the residues in IFABP to alanine and recomputing the DeltaDeltaG(binding). By postprocessing a single trajectory of the wild-type complex, the average unsigned error of our calculated DeltaDeltaG(binding) is below 1.5 kcal/mol for most of the alanine mutations of the noncharged residues (67% in total). To further investigate some particular mutants, three additional dynamical simulations of IFABP Arg126Ala, Arg106Ala, and Arg106Gln mutants were conducted. Recalculated binding free energies are well consistent with the experimental data. Moreover, the ambiguous role of Arg106 caused by the free energy change of the opposite sign when it is mutated to alanine and glutamine respectively is clarified both structurally and energetically. Typically, this can be attributed to the partial electrostatic compensation mainly from Arg56 and the obvious entropy gain in Arg106Ala mutant while not in Arg106Gln mutant. The presented structural model of IFABP-PA complex could be used to guide future studies.  相似文献   

4.
We synthesized Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu, the peptide contained in lactoferrin (Lf), to identify the angiotensin converting enzyme (ACE) inhibition. In an attempt to know the structure‐activity relationship of this peptide, we replaced Pro (the third amino acid residues from N‐terminal) or Val (the fourth amino acid residues from N‐terminal) with Ala (neutral amino acid), Glu (acidic amino acid) or Lys (basic amino acid) to produce six peptides. From the in vitro ACE inhibition (IC50) of these synthesized peptides, the original peptide (Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu) showed higher ACE inhibition than the replaced six peptides. Thus, replacement of Pro at the third amino acid residues or Val at the fourth position with Ala, Glu or Lys revealed the ACE inhibition to be lower than the original form of Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu. Otherwise, we added one peptide at the C‐terminal of Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu and found both products with an addition of Val (Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu‐Val) or Ile (Leu‐Arg‐Pro‐Val‐Ala‐Ala‐Glu‐Ile) showing a lower ACE inhibition than the original one. The ACE inhibitions produced by both replaced peptides were without significance. Also, deletion of the last peptide at the C‐terminal (Leu‐Arg‐Pro‐Val‐Ala‐Ala) failed to produce a marked change of ACE inhibition as compared to the original one. These results suggest that Pro and Val are essential in the peptide for inhibition of ACE activity.  相似文献   

5.
Absolute 18-crown-6 (18C6) affinities of five amino acids (AAs) are determined using guided ion beam tandem mass spectrometry techniques. The AAs examined in this work include glycine (Gly), alanine (Ala), lysine (Lys), histidine (His), and arginine (Arg). Theoretical electronic structure calculations are performed to determine stable geometries and energetics for neutral and protonated 18C6 and the AAs as well as the proton bound complexes comprised of these species, (AA)H(+)(18C6). The proton affinities (PAs) of Gly and Ala are lower than the PA of 18C6, whereas the PAs of Lys, His, and Arg exceed that of 18C6. Therefore, the collision-induced dissociation (CID) behavior of the (AA)H(+)(18C6) complexes differs markedly across these systems. CID of the complexes to Gly and Ala produces H(+)(18C6) as the dominant and lowest energy pathway. At elevated energies, H(+)(AA) was produced in competition with H(+)(18C6) as a result of the relatively favorable entropy change in the formation of H(+)(AA). In contrast, CID of the complexes to the protonated basic AAs results in the formation of H(+)(AA) as the only direct CID product. H(+)(18C6) was not observed, even at elevated energies, as a result of unfavorable enthalpy and entropy change associated with its formation. Excellent agreement between the measured and calculated (AA)H(+)-18C6 bond dissociation energies (BDEs) is found with M06 theory for all complexes except (His)H(+)(18C6), where theory overestimates the strength of binding. In contrast, B3LYP theory significantly underestimates the (AA)H(+)-18C6 BDEs in all cases. Among the basic AAs, Lys exhibits the highest binding affinity for 18C6, suggesting that the side chains of Lys residues are the preferred binding site for 18C6 complexation in peptides and proteins. Gly and Ala exhibit greater 18C6 binding affinities than Lys, suggesting that the N-terminal amino group provides another favorable binding site for 18C6. Trends in the 18C6 binding affinities among the five AAs examined here exhibit an inverse correlation with the polarizability and proton affinity of the AA. Therefore, the ability of the N-terminal amino group to compete for 18C6 complexation is best for Gly and should become increasing less favorable as the size of the side chain substituent increases.  相似文献   

6.
The mechanism of cis-trans isomerization of prolyl peptides by cyclophilin   总被引:3,自引:0,他引:3  
The mechanism of cis-trans isomerization of prolyl peptides catalyzed by cyclophilin (CyP) was studied computationally via molecular dynamics (MD) simulations of the transition state (TS) and the cis and trans forms of the ground state (GS), when bound to CyP and when free in aqueous solution. The MD simulations include four enzyme-bound species of tetrapeptide (Suc-Ala-XC([double bond]O)-NPro-Phe-pNA; X = Gly, Trp, Ala, and Leu). In water, the prolyl amide bond is favorably planar with the presence of conformers exhibiting +/-20 degrees twist of the C-N dihedral. In the active site a hydrogen bond between the cis-prolyl amide carbonyl O and the backbone amide N-H of Asn102 retains the 20 degrees twist of the C-N dihedral. The TS structure is characterized by a 90 degrees twist of the amide C-N bond and a more favorable interaction with Asn102 due to the shorter distance between Asn102(HN) and the amide carbonyl O. The conformational change of cis --> TS also involves pyramidalization of the amide N, which results in the formation of a hydrogen bond between the amide N and the guanidino group of Arg55. Both Asn102 and Arg55 are held in the same position in CyP.cis-isomer as in CyP.TS. In the ligand-free CyP the Arg55 guanidino group is highly disorganized and Asn102 is displaced 1 A from the position in the ligand-bound CyP. Thus, the organization of Arg55 and Asn102 occurs upon substrate binding. The geometrical complimentarity of the organized enzyme structure to the TS structure is a result of preferential binding of the proline N and the amide carbonyl of the TS compared to that of GS. However, the N-terminal part (Suc-Ala) becomes repositioned in the TS such that two hydrogen bonds disappear, one hydrogen bond appears and two other hydrogen bonds becomes weaker on the conversion of CyP.cis to CyP.TS. During this conversion, total hydrophobic contact between enzyme and the peptide is preserved. Thus, the interaction energies of GS and TS with enzyme are, as a whole, much alike. This does not support the contention that TS is bound more tightly than GS by K(m)/K(TS) = 10(6) in the cis --> trans reaction. Repositioning of the N-terminal part of the peptide on CyP.TS formation becomes more pronounced when the substrate X residue is changed from Gly < Trp < Ala < Leu. We propose that the larger turning of the N-terminus is responsible for the larger value of the experimentally observed Delta S(++) and Delta H(++), which sum up to little change in Delta G(++). The positioning of the Arg55 and the degree of 20 degrees twist of the amide C-N bond are considered as criteria for Near Attack Conformers (NACs) in cis-trans isomerization. NACs account for approximately 30% of the total GS populations of the cis-isomer. Similar NAC populations were observed with four different substrates. This is consistent with the insensitivity of enzymatic activity to the nature of the X residue. Also, the NAC population in CyP.trans-AAPF was comparable to that in CyP.cis-AAPF, in accord with similar experimentally measured rates of the cis --> trans and trans --> cis reaction in CyP. These NACs, found in CyP.cis and CyP.trans, resemble only one of the four possible TS configurations in the water reaction. The identity of this TS structure (syn/exo) is in accord with experimentally determined KIE values in the enzymatic reaction. However, the geometry of the active site was also complementary to another TS structure (anti/exo) that was not detected in the active site by the same KIE measurements, implying that the geometrical fitness of the TS cannot be a single determining factor for enzymatic reactions.  相似文献   

7.
The N-terminal 30-amino acid tail of histone H4, a nuclear protein, was studied as a model for the interaction of this protein with Ni(ii) ions. The behaviour of the ends-blocked Ac-SGRGKGGKGLGKGGA(15)K(16)R(17)H(18)R(19)KVLRDNIQGIT-Am fragment towards Ni(ii) was analyzed with multidimensional NMR (1D, 2D TOCSY, NOESY) and UV-Vis spectroscopy. As expected, the coordination involved the imidazolic nitrogen of the His(18) residue and the three deprotonated amidic nitrogens of the His(18), Arg(17) and Lys(16) residues, respectively. A model for the structure of the complex was calculated from the inter-residual NOEs recorded in 2D NOESY spectra. The structure obtained shows that the interaction with the metal is responsible for deep changes in the conformation of the peptide, blocking the side chain of Arg(17) and Lys(16) residues above the coordination plane. These structural modifications may be physiologically relevant to the mechanism of nickel carcinogenesis.  相似文献   

8.
Cation-pi interactions between amino acid side chains are increasingly being recognized as important structural and functional features of proteins and other biomolecules. Although these interactions have been found in static protein structures, they have not yet been detected in dynamic biomolecular systems. We determined, by (1)H NMR spectroscopic titrations, the energies of cation-pi interactions of the amino acid derivative AcLysOMe (1) with AcPheOEt (2) and with AcTyrOEt (3) in aqueous and three organic solvents. The interaction energy is substantial; it ranges from -2.1 to -3.4 kcal/mol and depends only slightly on the dielectric constant of the solvent. To assess the effects of auxiliary interactions and structural preorganization on formation of cation-pi interactions, we studied these interactions in the association of pentapeptides. Upon binding of the positively-charged peptide AcLysLysLysLysLysNH(2) (5) to the negatively-charged partner AcAspAspXAspAspNH(2) (6), in which X is Leu (6a), Tyr (6b), and Phe (6c), multiple interactions occur. Association of the two pentapeptides is dynamic. Free peptides and their complex are in fast exchange on the NMR time-scale, and 2D (1)H ROESY spectra of the complex of the two pentapeptides do not show intermolecular ROESY peaks. Perturbations of the chemical shifts indicated that the aromatic groups in peptides 6b and 6c were affected by the association with 5. The association constants K(A) for 5 with 6a and with 6b are nearly equal, (4.0 +/- 0.7) x 10(3) and (5.0 +/- 1.0) x 10(3) M(-)(1), respectively, while K(A) for 5 with 6c is larger, (8.3 +/- 1.3) x 10(3) M(-)(1). Molecular-dynamics (MD) simulations of the pentapeptide pairs confirmed that their association is dynamic and showed that cation-pi contacts between the two peptides are stereochemically possible. A transient complex between 5 and 6 with a prominent cation-pi interaction, obtained from MD simulations, was used as a template to design cyclic peptides C(X) featuring persistent cation-pi interactions. The cyclic peptide C(X) had a sequence in which X is Tyr, Phe, and Leu. The first two peptides do, but the third does not, contain the aromatic residue capable of interacting with a cationic Lys residue. This covalent construct offered conformational stability over the noncovalent complexes and allowed thorough studies by 2D NMR spectroscopy. Multiple conformations of the cyclic peptides C(Tyr) and C(Phe) are in slow exchange on the NMR time-scale. In one of these conformations, cation-pi interaction between Lys3 and Tyr9/Phe9 is clearly evident. Multiple NOEs between the side chains of residues 3 and 9 are observed; chemical-shift changes are consistent with the placement of the side chain of Lys3 over the aromatic ring. In contrast, the cyclic peptide C(Leu) showed no evidence for close approach of the side chains of Lys3 and Leu9. The cation-pi interaction persists in both DMSO and aqueous solvents. When the disulfide bond in the cyclic peptide C(Phe) was removed, the cation-pi interaction in the acyclic peptide AC(Phe) remained. To test the reliability of the pK(a) criterion for the existence of cation-pi interactions, we determined residue-specific pK(a) values of all four Lys side chains in all three cyclic peptides C(X). While NOE cross-peaks and perturbations of the chemical shifts clearly show the existence of the cation-pi interaction, pK(a) values of Lys3 in C(Tyr) and in C(Phe) differ only marginally from those values of other lysines in these dynamic peptides. Our experimental results with dynamic peptide systems highlight the role of cation-pi interactions in both intermolecular recognition at the protein-protein interface and intramolecular processes such as protein folding.  相似文献   

9.
Human voltage-gated potassium channel Kv1.3 is an important pharmacological target for the treatment of autoimmune and metabolic diseases. Increasing clinical demands stipulate an active search for efficient and selective Kv1.3 blockers. Here we present a new, reliable, and easy-to-use analytical system designed to seek for and study Kv1.3 ligands that bind to the extracellular vestibule of the K+-conducting pore. It is based on Escherichia coli spheroplasts with the hybrid protein KcsA-Kv1.3 embedded into the membrane, fluorescently labeled Kv1.3 blocker agitoxin-2, and confocal laser scanning microscopy as a detection method. This system is a powerful alternative to radioligand and patch–clamp techniques. It enables one to search for Kv1.3 ligands both among individual compounds and in complex mixtures, as well as to characterize their affinity to Kv1.3 channel using the “mix and read” mode. To demonstrate the potential of the system, we performed characterization of several known Kv1.3 ligands, tested nine spider venoms for the presence of Kv1.3 ligands, and conducted guided purification of a channel blocker from scorpion venom.
Figure
The scheme of a fluorescent analytical system designed to seek for and study Kv1.3 ligands that bind to the extracellular vestibule of the K+-conducting pore.  相似文献   

10.
KIA7, a peptide with a highly restricted set of amino acids (Lys, Ile, Ala, Gly and Tyr), adopts a specifically folded structure. Some amino acids, including Lys, Ile, Ala, Gly and His, form under the same putative prebiotic conditions, whereas different conditions are needed for producing Tyr, Phe and Trp. Herein, we report the 3D structure and conformational stability of the peptide KIA7H, which is composed of only Lys, Ile, Ala, Gly and His. When the imidazole group is neutral, this 20‐mer peptide adopts a four‐helix bundle with a specifically packed hydrophobic core. Therefore, one‐pot prebiotic proteins with well‐defined structures might have arisen early in chemical evolution. The Trp variant, KIA7W, was also studied. It adopts a 3D structure similar to that of KIA7H and its previously studied Tyr and Phe variants, but is remarkably more stable. When tested for ribonucleolytic activity, KIA7H, KIA7W and even short, unstructured peptides rich in His and Lys, in combination with Mg++, Mn++ or Ni++ (but not Cu++, Zn++ or EDTA) specifically cleave the single‐stranded region in an RNA stem–loop. This suggests that prebiotic peptide–divalent cation complexes with ribonucleolytic activity might have co‐inhabited the RNA world.  相似文献   

11.
The thermodynamic stability and oligomerization status of the tumor suppressor p53 tetramerization domain have been studied experimentally and theoretically. A series of hydrophilic mutations at Met-340 and Leu-344 of human p53 were designed to disrupt the hydrophobic dimer-dimer interface of the tetrameric oligomerization domain of p53. Meanfield calculations of the free energy of the solvated mutants as a function of interdimer distance were compared with experimental data on the thermal stability and oligomeric state [tetramer, dimer, or equilibrium mixture of both] of each mutant. The calculations predicted a decreasing stability and oligomeric state for the following amino acids at residue 340: Met [tetramer] > Ser Asp, His, Gin, > Glu, Lys [dimer], whereas the experimental results showed the following order: Met [tetramer] > Ser > Gln > His, Lys > Asp, Glu [dimers]. For residue 344, the calculated trend was Leu [tetramer] > Ala > Arg, Gln, Lys [dimer], and the experimental trend was Leu [tetramer] > Ala, Arg, Gln, Lys [dimer]. The discrepancy for the lysine side chain at residue 340 is attributed to the dual nature of lysine, both hydrophobic and charged. The incorrect prediction of stability of the mutant with Asp at residue 340 is attributed to the fact that within the meanfield approach, we use the wild-type backbone configuration for all mutants, but low melting temperatures suggest a softening of the α-helices at the dimer-dimer interface. This initial application of meanfield theory toward a protein-solvent system is encouraging for the application of the theoretical model to more complex systems.  相似文献   

12.
Electrostatic interactions dominate the structure and free energy of biomolecules. To obtain accurate free energies involving charged groups from molecular simulations, OPLS-AA parameters have been reoptimized using Monte Carlo free energy perturbation. New parameters fit a self-consistent, experimental set of hydration free energies for acetate (Asp), propionate (Glu), 4-methylimidazolium (Hip), n-butylammonium (Lys), and n-propylguanidinium (Arg), all resembling charged residue side chains, including beta-carbons. It is shown that OPLS-AA free energies depend critically on the type of water model, TIP4P or TIP3P; i.e., each water model requires specific water-charged molecule interaction potentials. New models (models 1 and 3) are thus described for both water models. Uncertainties in relative free energies of charged residues are approximately 2 kcal/mol with the new parameters, due to variations in system setup (MAEs of ca. 1 kcal/mol) and noise from simulations (ca. 1 kcal/mol). The latter error of approximately 1 kcal/mol contrasts MAEs from standard OPLS-AA of up to 13 kcal/mol for the entire series of charged residues or up to 5 kcal/mol for the cationic series Lys, Arg, and Hip. The new parameters can be used directly in molecular simulations with no modification of neutral residues needed and are envisioned to be particular important in simulations where charged residues change environment.  相似文献   

13.
The interaction mechanism of baicalein and lysozyme (Lys) has been characterized by fluorescence, synchronous fluorescence, ultraviolet-vis absorbance, and three-dimensional (3D) fluorescence. The structural characteristics of baicalein and Lys were probed, and their binding affinities were determined under different pH conditions (pH 7.4, 4.5, and 2.5). The results showed that the binding abilities of the drug to Lys increased under lower pH conditions (pH 4.5 and 2.5) due to the alterations of the protein secondary and tertiary structures or the structural change of baicalein. The effect of baicalein on the conformation of Lys was analyzed using UV, synchronous fluorescence and three-dimensional (3D) fluorescence under different pH conditions. These results indicate that the binding of baicalein to Lys causes apparent change in the secondary and tertiary structure of Lys. In the presence of Cu(2+), the decrease of the binding constant in buffer solution of pH 2.5 may result from the competition of the metal ion and baicalein binding to Lys. In addition, the presence of Cu(2+) increased the binding constants of baicalein-Lys complex under higher pH conditions (pH 7.4 and 4.5). The possible site of binding of baicalein to Lys has been proposed to explain these observations.  相似文献   

14.
ShK toxin is a cysteine‐rich 35‐residue protein ion‐channel ligand isolated from the sea anemone Stichodactyla helianthus. In this work, we studied the effect of inverting the side chain stereochemistry of individual Thr or Ile residues on the properties of the ShK protein. Molecular dynamics simulations were used to calculate the free energy cost of inverting the side‐chain stereochemistry of individual Thr or Ile residues. Guided by the computational results, we used chemical protein synthesis to prepare three ShK polypeptide chain analogues, each containing either an allo‐Thr or an allo‐Ile residue. The three allo‐Thr or allo‐Ile‐containing ShK polypeptides were able to fold into defined protein products, but with different folding propensities. Their relative thermal stabilities were measured and were consistent with the MD simulation data. Structures of the three ShK analogue proteins were determined by quasi‐racemic X‐ray crystallography and were similar to wild‐type ShK. All three ShK analogues retained ion‐channel blocking activity.  相似文献   

15.
We investigate the role of dynamics on adsorption of peptides to gold surfaces using all-atom molecular dynamics simulations in explicit solvent. We choose six homopolypeptides [Ala(10), Ser(10), Thr(10), Arg(10), Lys(10), and Gln(10)], for which experimental surface coverages are not correlated with amino acid level affinities for gold, with the idea that dynamic properties may also play a role. To assess dynamics we determine both conformational movement and flexibility of the peptide within a given conformation. Low conformational movement indicates stability of a given conformation and leads to less adsorption than homopolypeptides with faster conformational movement. Likewise, low flexibility within a given conformation also leads to less adsorption. Neither amino acid affinities nor dynamic considerations alone predict surface coverage; rather both quantities must be considered in peptide adsorption to gold surfaces.  相似文献   

16.
The DNA binding domain of transposon Tn916 integrase (INT‐DBD) binds to DNA target site by positioning the face of a three‐stranded antiparallel β‐sheet within the major groove. As the negatively charged DNA directly interacts with the positively charged residues (such as Arg and Lys) of INT‐DBD, the electrostatic interaction is expected to play an important role in the dynamical stability of the protein–DNA binding complex. In the current work, the combined use of quantum‐based polarized protein‐specific charge (PPC) for protein and polarized nucleic acid‐specific charge (PNC) for DNA were employed in molecular dynamics simulation to study the interaction dynamics between INT‐DBD and DNA. Our study shows that the protein–DNA structure is stabilized by polarization and the calculated protein–DNA binding free energy is in good agreement with the experimental data. Furthermore, our study revealed a positive correlation between the measured binding energy difference in alanine mutation and the occupancy of the corresponding residue's hydrogen bond. This correlation relation directly relates the contribution of a specific residue to protein–DNA binding energy to the strength of the hydrogen bond formed between the specific residue and DNA. © 2013 Wiley Periodicals, Inc.  相似文献   

17.
ShK toxin is a cysteine‐rich 35‐residue protein ion‐channel ligand isolated from the sea anemone Stichodactyla helianthus. In this work, we studied the effect of inverting the side chain stereochemistry of individual Thr or Ile residues on the properties of the ShK protein. Molecular dynamics simulations were used to calculate the free energy cost of inverting the side‐chain stereochemistry of individual Thr or Ile residues. Guided by the computational results, we used chemical protein synthesis to prepare three ShK polypeptide chain analogues, each containing either an allo‐Thr or an allo‐Ile residue. The three allo‐Thr or allo‐Ile‐containing ShK polypeptides were able to fold into defined protein products, but with different folding propensities. Their relative thermal stabilities were measured and were consistent with the MD simulation data. Structures of the three ShK analogue proteins were determined by quasi‐racemic X‐ray crystallography and were similar to wild‐type ShK. All three ShK analogues retained ion‐channel blocking activity.  相似文献   

18.
Matrix-assisted laser desorption/ionization (MALDI) mass spectrometry was used to study peptide-peptide interaction. The interaction was seen when 6-aza-2-thiothymine was used as a matrix (pH 5.4), but was disrupted with a more acidic matrix, alpha-cyano-4-hydroxycinnamic acid (pH 2.0). In the present study, we show that dynorphin, an opioid peptide, and five of its fragments that contain two adjacent basic residues (Arg6-Arg7), all interact noncovalently with peptides that contain two to five adjacent acidic residues (Asp or Glu). Two other nonrelated peptides containing two (Arg6-Arg7) or three (Arg1-Lys2-Arg3) adjacent basic amino acid residues were studied and exhibited the same behavior. However, peptides containing adjacent Lys or His did not form noncovalent complexes with acidic peptides. The noncovalent bonding was sufficiently stable that digestion with trypsin only cleaved Arg and Lys residues that were not involved in hydrogen bonding with the acidic residues. In an equimolar mixture of dynorphin, dynorphin fragments (containing the motif RR), and an acidic peptide (minigastrin), the acidic peptide preferentially complexed with dynorphin. If the concentration of minigastrin was increased 10 fold, noncovalent interaction was seen with dynorphin and all its fragments containing the motif RR. In the absence of dynorphin, minigastrin formed noncovalent complexes with all dynorphin fragments. These findings suggest that conformation, equilibrium, and concentration do play a role in the occurrence of peptide-peptide interaction. Observations from this study include: (1) ionic bonds were not disrupted by enzymatic digests, (2) conformation and concentration influenced complex formation, and (3) the complex did not form with fragments of dynorphin or unrelated peptides that did not contain the motifs RR or RKR, nor with a fragment of dynorphin where Arg7 was mutated to a phenylalanine residue. These findings strongly suggest that peptide-peptide interaction does occur, and can be studied by MALDI if near physiologic pH is maintained.  相似文献   

19.
视紫红质蛋白是一个跨膜蛋白, 视黄醛(RET)在该蛋白中的活性结合位点涉及到视觉过程机理, 与一些眼科疾病病理有关. 基于牛视紫红质蛋白1U19的蛋白质晶体结构数据, 采用密度泛函理论的B3LYP方法计算RET-Lys296残基与视黄醛分子周围半径为0.6 nm的空间范围30个氨基酸残基相互作用和结合能. 数值显示1U19蛋白中的残基Glu113、Glu181和Glu122是质子化的RET-Lys296残基的活性结合位点, 结合能分别为-333.38、-205.67和-194.56 kJ·mol-1. 这些氨基酸残基带有一个负电荷, 与质子化的RET-Lys296残基发生强烈的离子静电相互作用. 另外几个残基Ala292、Cys187、Phe293、Pro291以及Trp265等与质子化RET-Lys296残基也有相互吸引作用. 当RET-Lys296残基非质子化, 上述相互作用消失, 促使视黄醛分子与视蛋白分离. 研究发现残基Glu113和Glu181周围各自有一个结晶水分子通过双氢键形式起着稳定作用.  相似文献   

20.
Fourteen 42-residue polypeptides have been designed to identify reactive sites for the catalysis of the decarboxylation of oxaloacetate, a chemical transformation that proceeds through the formation of an imine intermediate. The sequences fold into helix-loop-helix motifs and dimerize to four-helix bundles. The catalytically active lysine residues were incorporated in several surface exposed positions, but also in positions characterised by hydrophobic properties to reduce their pKa values. The molecular environments of the Lys residues were systematically varied, to find which residues were able to stabilise and bind the imine intermediate in the decarboxylation reaction. A two-residue Arg-Lys site formed the main component of the reactive site of the helix-loop-helix dimer Decarb-K34_R33, which obeyed saturation kinetics in catalysing the reaction with a kcat/KM of 0.59 M-1S-1. The rate constant measured was nearly three orders of magnitude larger than the second-order rate constant of the butylamine-catalysed reaction (0.0011 M-1S-1), and four orders of magnitude larger than the pseudo first-order rate constant of the uncatalyzed reaction (1.3 x 10(-5) s(-1)). The sequence of Decarb-K34_R33 contained only a single lysine residue. It was flanked by an arginine in the preceding position in the sequence. A flanking Arg residue provided more efficient catalysis than a flanking Lys or Gln residue. Arginines in flanking positions in the helix, in positions four residues before or after the Lys in the sequence, are not as important in catalysis as the Arg of the Arg-Lys pair. The effect of pKa on the catalytic efficiency of the Lys residue in the decarboxylation reaction is well known. The identification of the role of the flanking Arg residue in catalysing decarboxylation, its optimal position, and the importance of conformational stability reported here sets the stage for developing a number of catalytic systems that depend on the formation of imine intermediates, but that lead to different reaction products.  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