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.
Collision-induced fragmentations of deprotonated maculatin 1 peptides provide significant sequencing information. When the peptide lacks those residues which can fragment through their alpha side chains (e.g. Thr, Ser, Glu and Gln in this study) the basic alpha and beta' backbone cleavages occur from the [Mbond;H](-) anion. When Thr, Ser, Glu and Gln are present, the ease of side-chain fragmentation of these residues is: Thr (loss of MeCHO) > Ser (CH(2)O) > Glu (H(2)O) > Gln (NH(3)). When one of more of these residues is (are) present, the alpha and beta' cleavages often occur from a fragment rather than the [Mbond;H](-) anion, e.g. for Thr, the [(Mbond;H)(-)bond;MeCHO](-) anion. These four residues also initiate gamma backbone cleavage reactions. The relative abundances of peaks resulting from gamma cleavage are Glu > Ser = Thr > Gln for maculatin 1 spectra. An unusual Gln19/Ile17 cyclisation/cleavage reaction occurs in maculatin spectra: the peptide [Mbond;H](-) anion must adopt a helical conformation in order for these two groups to interact. Analogous fragmentations have been reported previously in the negative ion spectra of the caerin 1 peptides.  相似文献   

2.
The metal-binding ability of human ubiquitin (hUb) towards a selection of biologically relevant metal ions and complexes has been probed. Different techniques have been used to obtain crystals suitable for crystallographic analysis. In the first type of experiments, crystals of hUb have been soaked in solutions containing copper(II) acetate and two metallodrugs, Zeise salt (K[PtCl(3)(η(2)-C(2)H(4))]·H(2)O) and cisplatin (cis-[PtCl(2)(NH(3))(2)]). The Zeise salt is used in a test for hepatitis, whereas cisplatin is one of the most powerful anticancer drugs in clinical use. The Zeise salt readily reacts with hUb crystals to afford an adduct with three platinum residues per protein molecule, Pt(3)-hUb. In contrast, copper(II) acetate and cisplatin were found to be unreactive for contact times up to one hour and to cause degradation of the hUb crystals for longer times. In the second type of experiments, hUb was cocrystallized with a solution of copper(II) or zinc(II) acetate or cisplatin. Zinc(II) acetate gives, at low metal-to-protein molar ratios (8:1), crystals containing one metal ion per three molecules of protein, Zn-hUb(3) (already reported in previous work), whereas at high metal-to-protein ratios (70:1) gives crystals containing three Zn(II) ions per protein molecule, Zn(3)-hUb. In contrast, once again, copper(II) acetate and cisplatin, even at low metal-to-protein ratios, do not give crystalline material. In the soaking experiment, the Zeise anion leads to simultaneous platination of His68, Met1, and Lys6. Present and previous results of cocrystallization experiments performed with Zn(II) and other Group 12 metal ions allow a comprehensive understanding of the metal-ion binding properties of hUb with His68 as the main anchoring site, followed by Met1 and carboxylic groups of Glu16, Glu18, Glu64, Asp21, and Asp32, to be reached. In the case of platinum, Lys6 can also be a binding site. The amount of bound metal ion, with respect to that of the protein, appears to be a relevant parameter influencing crystal packing.  相似文献   

3.
We report, for the first time, a detailed crystallographic study of the supramolecular arrangement for a set of zinc(II) Schiff base complexes containing the ligand 2,6-bis((E)-((2-(dimethylamino)ethyl)imino)methyl)-4-R-phenol], where R=methyl/tert-butyl/chloro. The supramolecular study acts as a pre-screening tool for selecting the compartmental ligand R of the Schiff base for effective binding with a targeted protein, bovine serum albumin (BSA). The most stable hexagonal arrangement of the complex [Zn − Me] (R=Me) stabilises the ligand with the highest FMO energy gap (ΔE=4.22 eV) and lowest number of conformations during binding with BSA. In contrast, formation of unstable 3D columnar vertebra for [Zn − Cl] (R=Cl) tend to activate the system with lowest FMO gap (3.75 eV) with highest spontaneity factor in molecular docking. Molecular docking analyses reported in terms of 2D LigPlot+ identified site A, a cleft of domains IB, IIIA and IIIB, as the most probable protein binding site of BSA. Arg144, Glu424, Ser428, Ile455 and Lys114 form the most probable interactions irrespective of the type of compartmental ligands R of the Schiff base whereas Arg185, Glu519, His145, Ile522 act as the differentiating residues with ΔG=−7.3 kcal mol−1.  相似文献   

4.
谷氨酰胺结合蛋白的分子动力学模拟和自由能计算   总被引:4,自引:0,他引:4  
胡建平  孙庭广  陈慰祖  王存新 《化学学报》2006,64(20):2079-2085
谷氨酰胺结合蛋白(Glutamine-binding protein, GlnBp)是大肠杆菌透性酶系统中一个细胞外液底物专一性结合蛋白, 对于细胞外液中谷氨酰胺(Gln)的运输和传递至关重要. 本文运用分子动力学(Molecular dynamics, MD)模拟采样, 考察了GlnBp关键残基与底物Gln之间的相互作用和GlnBp两条铰链的功能差别; 并采用MM-PBSA方法计算了GlnBp与底物Gln的结合自由能. 结果表明: Ph13, Phe50, Thr118和Ile69与底物Gln的范德华相互作用和Arg75, Thr70, Asp157, Gly68, Lys115, Ala67, His156与底物Gln的静电相互作用是结合Gln的主要推动力; 复合物的铰链区85~89柔性大, 对构象开合提供了结构基础; 而铰链区181~185柔性小, 其作用更多是在功能上把底物Gln限制在口袋中; 自由能预测值与实验值吻合. 本研究很好地解释了GlnBp结构与功能的关系, 为进一步了解GlnBp的开合及转运Gln的机制提供了重要的结构信息.  相似文献   

5.
In protein molecules each residue has a different ability to form contacts.In this paper,we calculated the number of contacts per residue and investigated the distribution of residue-residue contacts from 495 globular protein molecules using Contacts of Structural Units(CSU)software.It was found that the probability P(n)of amino acid residues having n pairs of contacts in all contacts fits Gaussian distribution very well.The distribution function of residue-residue contacts can be expressed as:P(n)=P_0+aexp[-b(n-n_c)~2].In our calculation,P_0=-0.06,α=11.4,b=-0.04 and n_c=9.0.According to distribution function,we found that those hydrophobic(H)residues including Leu,Val,Ile,Met,Phe,Tyr,Cys,and Trp residues have large values of the most probable number of contact n_c,and hydrophilic(P)residues including Ala,Gly,Thr, His,Glu,Gln,Asp,Asn,Lys,Ser,Arg,and Pro residues have the small ones.We also compare with Fauchere-Pliska hydrophobicity scale(FPH)and the most probable number of contact n_c for 20 amino acid residues,and find that there exists a linear relationship between Fauchere-Pliska hydrophobicity scale(FPH)and the most probable number of contact n_c, and it is expressed as:n_c=a+b×FPH,here α=8.87,and b=1.15.It is important to further explain protein folding and its stability from residue-residue contacts.  相似文献   

6.
Two peptide sequences from PARK9 Parkinson's disease gene, ProAspGluLysHisGluLeu, (P(1)D(2)E(3)K(4)H(5)E(6)L(7)) (1) and PheCysGlyAspGlyAlaAsnAspCysGly (F(1)C(2)G(3)D(4)G(5)A(6)N(7)D(8)C(9)G(10)) (2) were tested for Mn(II), Zn(II) and Ca(II) binding. The fragments are located from residues 1165 to 1171 and 1184 to 1193 in the PARK9 encoded protein. This protein can protect cells from poisoning of manganese, which is an environmental risk factor for a Parkinson's disease-like syndrome. Mono- and bi-dimensional NMR spectroscopy has been used to understand the details of metal binding sites at different pH values and at different ligand to metal molar ratios. Mn(II) and Zn(II) coordination with peptide (1) involves imidazole N(ε) or N(δ) of His(5) and carboxyl γ-O of Asp(2), Glu(3) and Glu(6) residues. Six donor atoms participate in Mn(II) binding resulting in a distorted octahedral geometry, possibly involving bidentate interaction of carboxyl groups; four donor atoms participate in Zn(II) binding resulting in a tetracoordinate geometry. Mn(II) and Zn(II) coordination involves the two cysteine residues with peptide (2); Mn(II) accepts additional ligand bonds from the carboxyl γ-O of Asp(4) and Asp(8) to complete the coordination sphere; the unoccupied sites may contain solvent molecules. The failure of Ca(II) ions to bind to either peptide (1) or (2) appears to result, under our conditions, from the absence of chelating properties in the chosen fragments.  相似文献   

7.
The X-ray crystal structure of the fluorescent antitumor sulfonamide carbonic anhydrase (CA, EC, 4.2.1.1) inhibitor (4-sulfamoylphenylethyl)thioureido fluorescein (1) in complex with the cytosolic isoform hCA II is reported, together with a modeling study of the adduct of 1 with the tumor-associated isoform hCA IX. Its binding to hCA II is similar to that of other benzesulfonamides, with the ionized sulfonamide coordinated to the Zn2+ ion within the enzyme active site, and also participating in a network of hydrogen bonds with residues Thr199 and Glu106. The scaffold of 1 did not establish polar interactions within the enzyme active site but made hydrophobic contacts (<4.5 A) with Gln92, Val121, Phe131, Val135, Leu198, Thr199, Thr200, and Pro202. The substituted 3-carboxy-amino-phenyl functionality was at van der Waals distance from Phe131, Gly132, and Val135. The bulky tricyclic fluorescein moiety was located at the rim of the active site, on the protein surface, and strongly interacted with the alpha-helix formed by residues Asp130-Val135. All these interactions were preserved in the hCA IX-1 adduct, but the carbonyl moiety of the fluorescein tail of 1 participates in a strong hydrogen bond with the guanidine moiety of Arg130, an amino acid characteristic of the hCA IX active site. This may account for the roughly 2 times higher affinity of 1 for hCA IX over hCA II and may explain why in vivo the compound specifically accumulates only in hypoxic tumors overexpressing CA IX and not in the normal tissues. The compound is in clinical studies as an imaging tool for acute hypoxic tumors.  相似文献   

8.
Supramolecular assembly of peptides and proteins into amyloid fibrils is of multifold interest, going from materials science to physiopathology. The binding of metal ions to amyloidogenic peptides is associated with several amyloid diseases, and amyloids with incorporated metal ions are of interest in nanotechnology. Understanding the mechanisms of amyloid formation and the role of metal ions can improve strategies toward the prevention of this process and enable potential applications in nanotechnology. Here, studies on Zn(II) binding to the amyloidogenic peptide Aβ11-28 are reported. Zn(II) modulates the Aβ11-28 aggregation, in terms of kinetics and fibril structures. Structural studies suggest that Aβ11-28 binds Zn(II) by amino acid residues Glu11 and His14 and that Zn(II) is rapidly exchanged between peptides. Structural and aggregation data indicate that Zn(II) binding induces the formation of the dimeric Zn(II)(1)(Aβ11-28)(2) species, which is the building block of fibrillar aggregates and explains why Zn(II) binding accelerates Aβ11-28 aggregation. Moreover, transient Zn(II) binding, even briefly, was enough to promote fibril formation, but the final structure resembled that of apo-Aβ11-28 amyloids. Also, seeding experiments, i.e., the addition of fibrillar Zn(II)(1)(Aβ11-28)(2) to the apo-Aβ11-28 peptide, induced aggregation but not propagation of the Zn(II)(1)(Aβ11-28)(2)-type fibrils. This can be explained by the dynamic Zn(II) binding between soluble and aggregated Aβ11-28. As a consequence, dynamic Zn(II) binding has a strong impact on the aggregation behavior of the Aβ11-28 peptide and might be a relevant and so far little regarded parameter in other systems of metal ions and amyloidogenic peptides.  相似文献   

9.
At a sufficiently high concentration of bovine oxyhemoglobin, the effect Zn(II) ions exert on its coagulation compares with that of Hg(II) ions. It is suggested that the center of preferential stoichiometric binding with Zn(II) ions in the protein is the sulfur atom of reactive SH groups. However, the binding of Zn(II) ions with groups other than thiolic equally affects the protein aggregation. Analysis of the pH dependence of the protein aggregation rate showed that the most likely alternative binding centers are histidine residues of the protein.  相似文献   

10.
The molecular chaperone αB‐crystallin, the major player in maintaining the transparency of the eye lens, preventing the aggregation of stress‐damaged and aging lens proteins from aggregation. In nonlenticular cells, it is involved in various neurological diseases, diabetes, and cancer. The role of some metal ions in the αB‐crystallin biology has been reported. Theoretical calculations have proposed that the coordination sites involving His101, His119, Lys121, His18 and Glu99 of human αB‐crystallin were the binding sites for divalent metal ions. Our previous mutagenesis study suggested that His18 rat lens αB‐crystallin is a crucial binding site for Cu(II) and Zn(II) in terms of chaperone‐like activity and structure. In this study mutant H119G of rat lens αB‐crystalin was cloned and expressed to investigate whether His119 is the coordination binding site. Copper and zinc at 1 mM concentration significantly increase the chaperone‐like activity in wild type αB‐crystalin, whereas zinc, copper and magnesium at 1 mM reduced the activity of H119G significantly. The results from chaperone‐like activity, ANS fluorescence measurement and Far‐and Near‐UV CD studies suggest that the replacement of His119 with Glycine resulted in a conformational and minor environmental changes that decrease chaperone‐like activity in the presence of divalent ions suggested that His119 was a crucial binding site for Cu(II) and Zn(II), which was similar to our previous study results of His18. Both results together suggest that His18 and His119 coordinates each other for the binding site of Cu(II) and Zn(II) in terms of improving the chaperone‐like activity and stability of crystallin/metal ion complex.  相似文献   

11.
The Hpn-like protein (Hpnl), a histidine- and glutamine-rich protein, is critical for Helicobacter pylori colonization in human gastric muscosa. In this study, the thermodynamic properties of Ni(II), Cu(II), Co(II), and Zn(II) toward Hpnl were studied by isothermal titration calorimetry (ITC). We found that Hpnl exhibits two independent binding sites for Ni(II) as opposed to one site for Cu(II), Co(II), and Zn(II). Protease digestion and chemical denaturation analysis further revealed that Ni(II) confers a higher stability upon Hpnl than other divalent metal ions. The potential Ni(II) binding sites are localized in the His-rich domain of Hpnl as confirmed by mutagenesis in combination with modification of histidine residues of the protein. We also demonstrated that the single mutants (H29A and H31A) and tetrameric mutant (H29-32A) cut nearly half of the binding capacity of Hpnl towards nickel ions, whereas other histidine residues (His30, 32, 38, 39, 40, and 41) are nonessential for nickel coordination. Escherichia coli cells that harbored H29A, H31A, and H29-32A mutant genes exhibited less tolerance toward high concentrations of extracellular nickel ions than those with the wild-type gene. Our combined data indicated that the conserved histidine residues, His29 and His31 in the His-rich domain of Hpnl, are critical for nickel binding, and such a binding is important for Hpnl protein to fulfill its biological functions.  相似文献   

12.
Scrambled ions have become the focus of recent investigations of peptide fragmentation. Here, an investigation of more than 390,000 high quality CID mass spectra is presented to explore the extent of scrambled ions in mass spectra and the possible fragmentation rules during scramble reactions. For the former, scrambled ions generally make up more than 10?% of mass spectra in number, although the abundances are less than 0.1 of the base peak. For the latter, relatively preferential re-opening sites were found for aliphatic residues Ala, Ile, Leu, and other residues such as Met, Gln, Ser, Phe, and Thr, whereas disfavored sites were found for basic residues Arg, Lys, and His, and Trp for both scrambled b and a ions. Similar preferential order in re-opening reaction was found in the reaction of losing internal residues when cleavage occurs at C-terminal side of 20 residues. However, when cleavage occurs at N-terminal side, Glu, Phe, and Trp become the most preferential sites. These results provide a deep insight into cleavage rules during scramble reactions for prediction of peptide mass spectra. Also, an additional investigation of whether scrambled ions could help discriminate false identifications from correct identifications was performed. Probing the number fraction of scrambled ions in falsely and correctly interpreted spectra and analyzing the correlation between scrambled ions and SEQUEST scores XCorr and Sp showed scrambled ions could at some extent help improve the discrimination in singly charged identifications, whereas no improvement was found for multiply charged results.  相似文献   

13.
We present mechanistic studies aimed at improving the understanding of the product ion formation rules in electron capture dissociation (ECD) of peptides and proteins in Fourier transform ion cyclotron resonance mass spectrometry. In particular, we attempted to quantify the recently reported general correlation of ECD product ion abundance (PIA) with amino acid hydrophobicity. The results obtained on a series of model H-RAAAAXAAAAK-OH peptides confirm a direct correlation of ECD PIA with X amino acid hydrophobicity and polarity. The correlation factor (R) exceeds 0.9 for 12 amino acids (Ile, Val, His, Asn, Asp, Glu, Gln, Ser, Thr, Gly, Cys, and Ala). The deviation of ECD PIA for seven outliers (Pro is not taken into consideration) is explained by their specific radical stabilization properties (Phe, Trp, Tyr, Met, and Leu) and amino acid basicity (Lys, Arg). Phosphorylation of Ser, Thr, and Tyr decreases the efficiency of ECD around phosphorylated residues, as expected. The systematic arrangement of amino acids reported here indicates a possible route toward development of a predictive model for quantitative electron capture/transfer dissociation tandem mass spectrometry, with possible applications in proteomics.  相似文献   

14.
15.
The melanocortin-1 receptor ( MC1R ) gene is known to play a major role in skin and hair pigmentation and to be highly polymorphic in Caucasians. This study was performed to investigate the relationships between MC1R gene polymorphisms and skin color in a large sample of French middle-aged Caucasian women. The codons 60 to 265 and the codon 294 of the MC1R gene were sequenced in 488 women. The skin color was measured on the inner side of the forearm using a spectrophotometric instrument. Fifteen variants were identified: Arg151Cys, Arg160Trp, Arg142His, Asp294His, Ile155Thr, Asp84Glu, Val60Leu, Val92Met, Arg163Gln, Ser83Pro, Thr95Met, Pro256Ser, Val265Ile, Ala166Ala and Gln233Gln. Women carrying Arg151Cys, Asp294His, Arg160Trp and Asp84Glu variants had a significantly higher reflectance in the red region, which indicates a lower level of functional melanin. This association was the most pronounced for women carrying Asp84Glu. In contrast, no significant difference was observed for other variants. Moreover, associations between MC1R polymorphisms and the risks of experiencing sunburn and of having freckles were found independently of skin color. Our findings support the hypothesis that MC1R polymorphisms do not necessarily alter the skin color but should sensitize the skin to UV-induced DNA damage.  相似文献   

16.
Chemical shift data from the BiomagResDataBank and conformational data derived from the protein data bank have been correlated in order to explore the conformational dependence of side chain (13)C resonance shifts. Consistent with predictions based on steric compression, upfield shifts for Cgamma resonances of Thr, Val, Ile, Leu, Met, Arg, Lys, Glu, and Gln residues correlate with both the number of heavy atom (nonproton) gamma-substituents and with gauche conformational orientations of gamma-substituents. The (13)C shift/conformation correlations are most apparent for Cgamma carbons but also can be observed at positions further from the backbone. Intraresidue steric conflict leads to a correlation between upfield-shifted side chain (13)C resonances and statistically lower probabilities in surveys of protein side chain conformation. Illustrative applications to the DNA pol lambda lyase domain and to dihydrofolate reductase are discussed. In the latter case, (13)C shift analysis indicates that the conformation of the remote residue V119 on the betaF-betaG loop is correlated with the redox state of the bound pyridine nucleotide cofactor, providing one basis for discrimination between substrate and product. It is anticipated that (13)C shift data for protein sidechains can provide a useful basis for the analysis of conformational changes even in large, deuterated proteins. Additionally, the large dependence of the leucine methyl shift difference, deltaCdelta1-deltaCdelta2, on both chi1 and chi2 is sufficient to allow this parameter to be used as a restraint in structure calculations if stereospecific assignment data are available.  相似文献   

17.
Protein-ligand interactions may lead to the formation of multiple molecular complexes in dynamic exchange, affecting the kinetic and thermodynamic characteristics of the binding equilibrium. We followed the dissociation kinetics of the transient and specific complex of an antithrombotic peptide N-acetyl-Asp(55)-Phe-Glu-Glu-Ile-Pro(60)-Glu-Glu-Tyr-Leu-Gln(65) with human prothrombin by use of (15)N NMR relaxation dispersion spectroscopy of the peptide. Every one of the five (15)N-labeled adjacent residues of the peptide exhibited apparently different kinetic exchange and relaxation behaviors, which were especially evident at different concentrations of prothrombin. Binding-induced (15)N relaxation dispersion of residues Phe(56), Glu(57), Glu(58), and Ile(59) can be fitted phenomenologically to a two-site on-and-off exchange mechanism with physically feasible relaxation and kinetic parameters obtained for residues Phe(56), Glu(58), and Ile(59), independent of the prothrombin concentration. The apparent kinetic parameters of Glu(57) show some dependence on the concentration of prothrombin and the extracted transverse relaxation rate for Glu(57) in the bound state was severalfold higher than that expected for a protein-peptide complex with a size of approximately 72 kDa. In addition, the equilibrium population of the bound peptide obtained for Glu(57) was inconsistent with those for Phe(56), Glu(58), and Ile(59) and with the prothrombin/peptide ratios used in the experiments. These discrepancies can be explained by the presence of two conformations for the peptide-protein complex exchanging at a rate of approximately 100 s(-)(1). In all, our study shows that fast dissociation of protein-peptide complexes can be studied quantitatively using peptide (15)N NMR relaxation dispersion measurements without a precise knowledge of the peptide and protein concentrations. In addition, protein titration was found to improve the accuracy of quantitative analysis and may make it possible to determine the rate of conformational changes within the protein-peptide complex.  相似文献   

18.
From the standpoint of protein dynamics and metalloprotein design, it is interesting to create an artificial protein which induces structural change and regulates its function by metal-ion binding. We engineered a novel protein, "Antennafinger (Ant-F)", whose structure and function can be controlled with Zn(II), by introducing the consensus sequence of a Cys(2)His(2)-type zinc finger protein into a non-metalloprotein scaffold, an Antennapedia homeodomain mutant (Ant-wt), selected using a motif-searching system. The circular dichroism studies demonstrate that Ant-F has secondary structures similar to Ant-wt and also changes its conformation due to Zn(II)-binding. The optical absorption spectra of the Co(II) complexes of Ant-F and its derivative proteins suggest that the geometry of the metal center of holo-Ant-F is tetrahedral and that the mutated Cys(2)His(2) residues are involved in the complex formation. In addition, the gel mobility shift assay reveals that the DNA binding activity of Ant-F can be regulated through Zn(II)-induced structural alteration. These results provide valuable information about the dynamic properties of proteins and a novel concept for metalloprotein design.  相似文献   

19.
Nature's use of biomineralization polypeptides to control and modulate the growth of biogenic minerals is an important process that, if properly understood, could have significant implications for designing and creating new inorganic-based materials. Although the sequences for a number of biomineralization proteins exist, very little is known about the participation of specific amino acids in the mineral modulation process. In this letter, we investigate the impact of global Asp --> Asn and Glu --> Gln substitutions on two mollusk shell nacre polypeptides, AP7N and n16N. We find that these global substitutions, which remove all anionic Ca(II) binding sites, abolish the expected in vitro mineralization activities associated with each native polypeptide. In addition, the ability of substituted peptides to form complexes with both Ca(II) and Ca(II) metal ion analogs is also abolished. However, some unexpected effects were noted. First, the Asp --> Asn, Glu --> Gln substituted n16N polypeptide is observed to self-assemble and form biofilms or coatings that appear to mineralize in vitro. Second, both polypeptides are structurally affected by these substitutions, with Asp --> Asn substituted AP7N transforming to an alpha helix and Asp --> Asn, Glu --> Gln substituted n16N transforming to a more unfolded random-coil-like structure. We find that the participation of Asp and Glu residues is crucial to the inherent mineralization activities and conformations of AP7N and n16N polypeptides. Surprisingly, we find that the replacement of anionic residues within biomineralization polypeptides such as n16N still permits mineral modulation, but in a different form that now involves peptide self-association and biofilm formation.  相似文献   

20.
The role of the second shell in the process of metal binding and selectivity in metalloproteins has been elucidated by combining Protein Data Bank (PDB) surveys of Mg, Mn, Ca, and Zn binding sites with density functional theory/continuum dielectric methods (DFT/CDM). Peptide backbone groups were found to be the most common second-shell ligand in Mg, Mn, Ca, and Zn binding sites, followed (in decreasing order) by Asp/Glu, Lys/Arg, Asn/Gln, and Ser/Thr side chains. Aromatic oxygen- or nitrogen-containing side chains (Tyr, His, and Trp) and sulfur-containing side chains (Cys and Met) are seldom found in the second coordination layer. The backbone and Asn/Gln side chain are ubiquitous in the metal second coordination layer as their carbonyl oxygen and amide hydrogen can act as a hydrogen-bond acceptor and donor, respectively, and can therefore partner practically every first-shell ligand. The second most common outer-shell ligand, Asp/Glu, predominantly hydrogen bonds to a metal-bound water or Zn-bound histidine and polarizes the H-O or H-N bond. In certain cases, a second-shell Asp/Glu could affect the protonation state of the metal ligand. It could also energetically stabilize a positively charged metal complex more than a neutral ligand such as the backbone and Asn/Gln side chain. As for the first shell, the second shell is predicted to contribute to the metal selectivity of the binding site by discriminating between metal cations of different ionic radii and coordination geometries. The first-shell-second-shell interaction energies decay rapidly with increasing solvent exposure of the metal binding site. They are less favorable but are of the same order of magnitude as compared to the respective metal-first-shell interaction energies. Altogether, the results indicate that the structure and properties of the second shell are dictated by those of the first layer. The outer shell is apparently designed to stabilize/protect the inner-shell and complement/enhance its properties.  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