共查询到20条相似文献,搜索用时 0 毫秒
1.
Del Vecchio P Carullo P Barone G Pagano B Graziano G Iannetti A Acquaviva R Leonardi A Formisano S 《Proteins》2008,70(3):748-760
2.
The binding of four epitope-related peptides and three library-derived, epitope-unrelated peptides of different lengths (10-14 amino acids) and sequence by anti-p24 (HIV-1) monoclonal antibody CB4-1 and its Fab fragment was studied by isothermal titration calorimetry. The binding constants K(A) at 25 degrees C vary between 5.1 x 10(7) M (-1) for the strongest and 1.4 x 10(5) M (-1) for the weakest binder. For each of the peptides complex formation is enthalpically driven and connected with unfavorable entropic contributions; however, the ratio of enthalpy and entropy contributions to deltaG(0) differs markedly for the individual peptides. A plot of -deltaH(0) vs -TdeltaS(0) shows a linear correlation of the data for a wide variety of experimental conditions as expected for a process with deltaC(p) much larger than deltaS(0). The dissimilarity of deltaC(p) and deltaS(0) also explains why deltaH(0) and TdeltaS(0) show similar temperature dependences resulting in relatively small changes of deltaG(0) with temperature. The heat capacity changes deltaC(p) upon antibody-peptide complex formation determined for three selected peptides vary only in a small range, indicating basic thermodynamic similarity despite different key residues interacting in the complexes. Furthermore, the comparison of van't Hoff and calorimetric enthalpies point to a non-two-state binding mechanism. Protonation effects were excluded by measurements in buffers of different ionization enthalpies. Differences in the solution conformation of the peptides as demonstrated by circular dichroic measurements do not explain different binding affinities of the peptides; specifically a high helix content in solution is not essential for high binding affinity despite the helical epitope conformation in the crystal structure of p24. 相似文献
3.
Rongjin Guan Sriram Aiyer Marie L. Cote Rong Xiao Mei Jiang Thomas B. Acton Monica J. Roth Gaetano T. Montelione 《Proteins》2017,85(4):647-656
The retroviral integrase (IN) carries out the integration of a dsDNA copy of the viral genome into the host DNA, an essential step for viral replication. All IN proteins have three general domains, the N‐terminal domain (NTD), the catalytic core domain, and the C‐terminal domain. The NTD includes an HHCC zinc finger‐like motif, which is conserved in all retroviral IN proteins. Two crystal structures of Moloney murine leukemia virus (M‐MuLV) IN N‐terminal region (NTR) constructs that both include an N‐terminal extension domain (NED, residues 1–44) and an HHCC zinc‐finger NTD (residues 45–105), in two crystal forms are reported. The structures of IN NTR constructs encoding residues 1–105 (NTR1–105) and 8–105 (NTR8–105) were determined at 2.7 and 2.15 Å resolution, respectively and belong to different space groups. While both crystal forms have similar protomer structures, NTR1–105 packs as a dimer and NTR8–105 packs as a tetramer in the asymmetric unit. The structure of the NED consists of three anti‐parallel β‐strands and an α‐helix, similar to the NED of prototype foamy virus (PFV) IN. These three β‐strands form an extended β‐sheet with another β‐strand in the HHCC Zn2+ binding domain, which is a unique structural feature for the M‐MuLV IN. The HHCC Zn2+ binding domain structure is similar to that in HIV and PFV INs, with variations within the loop regions. Differences between the PFV and MLV IN NEDs localize at regions identified to interact with the PFV LTR and are compared with established biochemical and virological data for M‐MuLV. Proteins 2017; 85:647–656. © 2016 Wiley Periodicals, Inc. 相似文献
4.
Strel'tsov S. A. Mikheikin A. L. Grokhovsky S. L. Oleinikov V. A. Kudelina I. A. Zhuze A. L. 《Molecular Biology》2002,36(5):736-753
Interaction of topotecan (TPT) with synthetic double-stranded polydeoxyribonucleotides has been studied in solutions of low ionic strength at pH 6.8 by linear flow dichroism (LD), circular dichroism (CD), UV-Vis absorption, and Raman spectroscopy. The complexes of TPT with poly(dG-dC)·poly(dG-dC), poly(dG)·poly(dC), poly(dA-dC)·poly(dG-dT), and poly(dA)·poly(dT), as well as complexes of TPT with calf thymus DNA and coliphage T4 DNA studied by us previously, have been shown to have negative LD in the long-wavelength absorption band of TPT, whereas the complex of TPT with poly(dA-dT)·poly(dA-dT) has positive LD in this absorption band of TPT. Thus, there are two different types of TPT complex with the polymers. TPT has been established to bind preferably to GC base pairs because its affinity to the polymers of different composition decreases in the following order: poly(dG-dC)·poly(dG-dC) > poly(dG)·poly(dC) > poly(dA-dC)·poly(dG-dT) > poly(dA)·poly(dT). The presence of DNA has been shown to shift the monomer–dimer equilibrium in TPT solutions toward dimer formation. Several duplexes of the synthetic polynucleotides bound together by bridges of TPT dimers may participate in the formation of the studied type of TPT–polynucleotide complex. Molecular models of TPT complex with linear and circular supercoiled DNAs and with deoxyguanosine have been considered. TPT (and presumably the whole camptothecin family) proved to represent a new class of DNA-specific ligands whose biological action is associated with formation of dimeric bridges between two DNA duplexes. 相似文献
5.
To evaluate the biological preference of chiral drug candidates for molecular target DNA, new potential metal‐based chemotherapeutic agents 1 , 1a , 1b , 2 , 2a , 2b , 3 , 3a , 3b of late 3d transition metals Ni(II), Cu(II), and Zn(II), respectively, derived from (R)‐ and (S)‐2‐amino‐2‐phenylethanol with CH2 CH2 linker were synthesized and thoroughly characterized. Interaction studies of 1 , 1a , 1b , 2 , 2a , 2b , 3 , 3a , 3b with calf thymus DNA in Tris buffer were studied by electronic absorption titrations, luminescence titrations, cyclic voltammetry, and circular dichroism. The results reveal that the extent of DNA binding of R‐enantiomer of copper 1a was highest in comparison to rest of the complexes via electrostatic interaction mode. The nuclease activity of 1a , 1b with supercoiled pBR322 DNA was further examined by gel electrophoresis, which reveals that complex 1a exhibits a remarkable DNA cleavage activity (concentration dependent) with pBR322DNA, and the cleavage activity of both enantiomers of complex 1 was significantly enhanced in the presence of activators. The activating efficiency follows the order Asc > H2O2 > MPA for 1a , and reverse order was observed for 1b , because of the differences in enantioselectivity and conformation. Further, it was observed that cleavage reaction involves singlet oxygen species and superoxide radicals via oxidative cleavage mechanism. In addition, complex 1a exhibits significant inhibitory effects on the topoisomerase II (topo II) activity at a very low concentration ∼24 μM, which suggest that complex 1a is indeed catalytic inhibitor or (poison) of human topo II. Chirality 2011 © 2011 Wiley‐Liss, Inc. 相似文献
6.
Welfle K Misselwitz R Sabat R Volk HD Schneider-Mergener J Reineke U Welfle H 《Journal of molecular recognition : JMR》2001,14(2):89-98
The mechanism of recognition of proteins and peptides by antibodies and the factors determining binding affinity and specificity are mediated by essentially the same features. However, additional effects of the usually unfolded and flexible solution structure of peptide ligands have to be considered. In an earlier study we designed and optimized six peptides (pepI to pepVI) mimicking the discontinuous binding site of interleukin-10 for the anti-interleukin-10 monoclonal antibody (mab) CB/RS/1. Three of them were selected for analysis of their solution conformation by circular dichroism measurements. The peptides differ in the content of alpha-helices and in the inducibility of helical secondary structures by trifluoroethanol. These properties, however, do not correlate with the binding affinity. PepVI, a 32-mer cyclic epitope mimic, has the highest affinity to mab CB/RS/1 identified to date. CD difference spectroscopy suggests an increase of the alpha-helix content of pepVI with complex formation. Binding of pepVI to mab CB/RS/1 is characterized by a large negative, favorable binding enthalpy and a smaller unfavorable loss of entropy (DeltaH degrees = -16.4 kcal x mol(-1), TDeltaS degrees = -6.9 kcal x mol(-1)) resulting in DeltaG degrees = -9.5 kcal x mol(-1) at 25 degrees C as determined by isothermal titration calorimetry. Binding of pepVI is enthalpically driven over the entire temperature range studied (10-35 degrees C). Complex formation is not accompanied by proton uptake or release. A negative heat capacity change DeltaC(p) of -0.354 kcal x mol(-1) x K(-1) was determined from the temperature dependence of DeltaH degrees. The selection of protein mimics with the observed thermodynamic properties is promoted by the applied identification and iterative optimization procedure. 相似文献
7.
R. Tyler-Cross M. Sobel D. Marques R. B. Harris 《Protein science : a publication of the Protein Society》1994,3(4):620-627
The serine proteinase inhibitor antithrombin III (ATIII) is a key regulatory protein of intrinsic blood coagulation. ATIII attains its full biological activity only upon binding polysulfated oligosaccharides, such as heparin. A series of synthetic peptides have been prepared based on the proposed heparin binding regions of ATIII and their ability to bind heparin has been assessed by CD spectrometry, by isothermal titration calorimetry, and by the ability of the peptides to compete with ATIII for binding heparin in a factor Xa procoagulant enzyme assay. Peptide F123-G148, which encompasses both the purported high-affinity pentasaccharide binding region and an adjacent, C-terminally directed segment of ATIII, was found to bind heparin with good affinity, but amino-terminal truncations of this sequence, including L130-G148 and K136-G148 displayed attenuated heparin binding activities. In fact, K136-G148 appears to encompass only a low-affinity heparin binding site. In contrast, peptides based solely on the high-affinity binding site (K121-A134) displayed much higher affinities for heparin. By CD spectrometry, these high-affinity peptides are chiefly random coil in nature, but low microM concentrations of heparin induce significant alpha-helix conformation. K121-A134 also effectively competes with ATIII for binding heparin. Thus, through the use of synthetic peptides that encompass part, if not all, of the heparin binding site(s) within ATIII, we have further elucidated the structure-function relations of heparin-ATIII interactions. 相似文献
8.
Previous isothermal titration calorimetry (ITC) and Förster resonance energy transfer studies demonstrated that Escherichia coli HUαβ binds nonspecifically to duplex DNA in three different binding modes: a tighter-binding 34-bp mode that interacts with DNA in large (> 34 bp) gaps between bound proteins, reversibly bending it by 140o and thereby increasing its flexibility, and two weaker, modestly cooperative small site-size modes (10 bp and 6 bp) that are useful for filling gaps between bound proteins shorter than 34 bp. Here we use ITC to determine the thermodynamics of these binding modes as a function of salt concentration, and we deduce that DNA in the 34-bp mode is bent around—but not wrapped on—the body of HU, in contrast to specific binding of integration host factor. Analyses of binding isotherms (8-bp, 15-bp, and 34-bp DNA) and initial binding heats (34-bp, 38-bp, and 160-bp DNA) reveal that all three modes have similar log-log salt concentration derivatives of the binding constants (Ski) even though their binding site sizes differ greatly; the most probable values of Ski on 34-bp DNA or larger DNA are − 7.5 ± 0.5. From the similarity of Ski values, we conclude that the binding interfaces of all three modes involve the same region of the arms and saddle of HU. All modes are entropy-driven, as expected for nonspecific binding driven by the polyelectrolyte effect. The bent DNA 34-bp mode is most endothermic, presumably because of the cost of HU-induced DNA bending, while the 6-bp mode is modestly exothermic at all salt concentrations examined. Structural models consistent with the observed Ski values are proposed. 相似文献
9.
Pawel Drozdzal Miroslaw Gilski Ryszard Kierzek Lechoslaw Lomozik Mariusz Jaskolski 《Acta Crystallographica. Section D, Structural Biology》2013,69(6):1180-1190
X‐ray crystal structures of the spermine4+ form of the Z‐DNA duplex with the self‐complementary d(CG)3 sequence in complexes with Mn2+ and Zn2+ cations have been determined at the ultrahigh resolutions of 0.75 and 0.85 Å, respectively. Stereochemical restraints were only used for the sperminium cation (in both structures) and for nucleotides with dual conformation in the Zn2+ complex. The Mn2+ and Zn2+ cations at the major site, designated M2+(1), bind at the N7 position of G6 by direct coordination. The coordination geometry of this site was octahedral, with complete hydration shells. An additional Zn2+(2) cation was bis‐coordinated in a tetrahedral fashion by the N7 atoms of G10 and G12 from a symmetry‐related molecule. The coordination distances of Zn2+(1) and Zn2+(2) to the O6 atom of the guanine residues were 3.613 (6) and 3.258 (5) Å, respectively. Moreover, a chloride ion was also identified in the coordination sphere of Zn2+(2). Alternate conformations were observed in the Z‐DNA–Zn2+ structure not only at internucleotide linkages but also at the terminal C3′—OH group of G12. The conformation of the sperminium chain in the Z‐DNA–Mn2+ complex is similar to the spermine4+ conformation in analogous Z‐DNA–Mg2+ structures. In the Z‐DNA–Zn2+ complex the sperminium cation is disordered and partially invisible in electron‐density maps. In the Z‐DNA–Zn2+ complex the sperminium cation only interacts with the phosphate groups of the Z‐DNA molecules, while in the Z‐DNA–Mn2+ structure it forms hydrogen bonds to both the phosphate groups and DNA bases. 相似文献
10.
Interaction of clinically important human DNA topoisomerase I poison, topotecan, with double-stranded DNA 总被引:2,自引:0,他引:2
Streltsov S Oleinikov V Ermishov M Mochalov K Sukhanova A Nechipurenko Y Grokhovsky S Zhuze A Pluot M Nabiev I 《Biopolymers》2003,72(6):442-454
Topotecan (TPT), a water-soluble derivative of camptothecin, is a potent antitumor poison of human DNA topoisomerase I (top1) that stabilizes the cleavage complex between the enzyme and DNA. The role of the recently discovered TPT affinity to DNA remains to be defined. The aim of this work is to clarify the molecular mechanisms of the TPT-DNA interaction and to propose the models of TPT-DNA complexes in solution in the absence of top1. It is shown that TPT molecules form dimers with a dimerization constant of (4.0 +/- 0.7) x 10(3) M(-1) and the presence of DNA provokes more than a 400-fold increase of the effective dimerization constant. Flow linear dichroism spectroscopy accompanied by circular dichroism, fluorescence, and surface-enhanced Raman scattering experiments provide evidence that TPT dimers are able to bind DNA by bridging different DNA molecules or distant DNA structural domains. This effect may provoke modification of the intrinsic geometry of the cruciform DNA structures, leading to the appearance of new crossover points that serve as the sites of the top1 loading position. The data presume the hypothesis of TPT-mediated modulation of top1-DNA recognition before ternary complex formation. 相似文献
11.
D. Wyrzykowski A. Tesmar D. Jacewicz J. Pranczk L. Chmurzyński 《Journal of molecular recognition : JMR》2014,27(12):722-726
The isothermal titration calorimetry (ITC) technique supported by potentiometric titration data was used to study the interaction of zinc ions with pH buffer substances, namely 2‐(N‐morpholino)ethanesulfonic acid (Mes), piperazine‐N,N′‐bis(2‐ethanesulfonic acid) (Pipes), and dimethylarsenic acid (Caco). The displacement ITC titration method with nitrilotriacetic acid as a strong, competitive ligand was applied to determine conditional–independent thermodynamic parameters for the binding of Zn(II) to Mes, Pipes, and Caco. Furthermore, the relationship between the proposed coordination mode of the buffers and the binding enthalpy has been discussed. Copyright © 2014 John Wiley & Sons, Ltd. 相似文献
12.
13.
锌(Zn)是生命代谢中重要的微量金属元素,但锌过量会对细胞造成毒害作用。细菌通过一些特有的机制来解除重金属离子对它们的毒害。该文重点介绍了细菌对高浓度Zn2+的抗性机制,主要包括外排机制(RND蛋白家族、CDF蛋白家族和P-型ATPase)、螯合机制和外排后结合机制。通过这些机制细菌能有效控制胞内Zn2+浓度,保护其不受过量Zn2+的毒害,但抗性机制往往不依赖单一的抗性系统,而是多种系统协调作用的结果。细菌Zn2+抗性机制的研究将有助于进一步揭示生物是如何应对高浓度金属离子的胁迫及相应的适应性规律。 相似文献
14.
A major goal in ligand and drug design is the optimization of the binding affinity of selected lead molecules. However, the binding affinity is defined by the free energy of binding, which, in turn, is determined by the enthalpy and entropy changes. Because the binding enthalpy is the term that predominantly reflects the strength of the interactions of the ligand with its target relative to those with the solvent, it is desirable to develop ways of predicting enthalpy changes from structural considerations. The application of structure/enthalpy correlations derived from protein stability data has yielded inconsistent results when applied to small ligands of pharmaceutical interest (MW < 800). Here we present a first attempt at an empirical parameterization of the binding enthalpy for small ligands in terms of structural information. We find that at least three terms need to be considered: (1) the intrinsic enthalpy change that reflects the nature of the interactions between ligand, target, and solvent; (2) the enthalpy associated with any possible conformational change in the protein or ligand upon binding; and, (3) the enthalpy associated with protonation/deprotonation events, if present. As in the case of protein stability, the intrinsic binding enthalpy scales with changes in solvent accessible surface areas. However, an accurate estimation of the intrinsic binding enthalpy requires explicit consideration of long-lived water molecules at the binding interface. The best statistical structure/enthalpy correlation is obtained when buried water molecules within 5-7 A of the ligand are included in the calculations. For all seven protein systems considered (HIV-1 protease, dihydrodipicolinate reductase, Rnase T1, streptavidin, pp60c-Src SH2 domain, Hsp90 molecular chaperone, and bovine beta-trypsin) the binding enthalpy of 25 small molecular weight peptide and nonpeptide ligands can be accounted for with a standard error of +/-0.3 kcal x mol(-1). 相似文献
15.
Zhenyao Luo Jacqueline R. Morey Christopher A. McDevitt Botjan Kobe 《Acta Crystallographica. Section F, Structural Biology Communications》2015,71(12):1459-1464
Zn2+ is an essential nutrient for all known forms of life. In the major human pathogen Streptococcus pneumoniae, the acquisition of Zn2+ is facilitated by two Zn2+‐specific solute‐binding proteins: AdcA and AdcAII. To date, there has been a paucity of structural information on AdcA, which has hindered a deeper understanding of the mechanism underlying pneumococcal Zn2+ acquisition. Native AdcA consists of two domains: an N‐terminal ZnuA domain and a C‐terminal ZinT domain. In this study, the ZnuA domain of AdcA was crystallized. The initial crystals of the ZnuA‐domain protein were obtained using dried seaweed as a heterogeneous nucleating agent. No crystals were obtained in the absence of the heterogeneous nucleating agent. These initial crystals were subsequently used as seeds to produce diffraction‐quality crystals. The crystals diffracted to 2.03 Å resolution and had the symmetry of space group P1. This study demonstrates the utility of heterogeneous nucleation. The solution of the crystal structures will lead to further understanding of Zn2+ acquisition by S. pneumoniae. 相似文献
16.
Sunghark Kwon Yuichi Nishitani Satoshi Watanabe Yoshinori Hirao Tadayuki Imanaka Tamotsu Kanai Haruyuki Atomi Kunio Miki 《Proteins》2016,84(9):1321-1327
A [NiFe] hydrogenase maturation protease HybD from Thermococcus kodakarensis KOD1 (TkHybD) is involved in the cleavage of the C‐terminal residues of [NiFe] hydrogenase large subunits by Ni recognition. Here, we report the crystal structure of TkHybD at 1.82 Å resolution to better understand this process. TkHybD exhibits an α/β/α sandwich fold with conserved residues responsible for the Ni recognition. Comparisons of TkHybD with homologous proteins also reveal that they share a common overall architecture, suggesting that they have similar catalytic functions. Our results including metal binding site prediction provide insight into the substrate recognition and catalysis mechanism of TkHybD. Proteins 2016; 84:1321–1327. © 2016 Wiley Periodicals, Inc. 相似文献
17.
The modes of binding of 5′‐[4‐(aminoiminomethyl)phenyl]‐[2,2′‐Bifuran]‐5‐carboximidamide (DB832) to multi‐stranded DNAs: human telomere quadruplex, monomolecular R‐triplex, pyr/pur/pyr triplex consisting of 12 T*(T·A) triplets, and DNA double helical hairpin were studied. The optical adsorption of the ligand was used for monitoring the binding and for determination of the association constants and the numbers of binding sites. CD spectra of DB832 complexes with the oligonucleotides and the data on the energy transfer from DNA bases to the bound DB832 assisted in elucidating the binding modes. The affinity of DB832 to the studied multi‐stranded DNAs was found to be greater (Kass ≈ 107M?1) than to the duplex DNA (Kass ≈ 2 × 105M?1). A considerable stabilizing effect of DB832 binding on R‐triplex conformation was detected. The nature of the ligand tight binding differed for the studied multi‐stranded DNA depending on their specific conformational features: recombination‐type R‐triplex demonstrated the highest affinity for DB832 groove binding, while pyr/pur/pyr TTA triplex favored DB832 intercalation at the end stacking contacts and the human telomere quadruplex d[AG3(T2AG3)3] accommodated the ligand in a capping mode. Additionally, the pyr/pur/pyr TTA triplex and d[AG3(T2AG3)3] quadruplex bound DB832 into their grooves, though with a markedly lesser affinity. DB832 may be useful for discrimination of the multi‐sranded DNA conformations and for R‐triplex stabilization. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 8–20, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com 相似文献
18.
Annfrid Sivertsen Bjørn Olav Brandsdal John Sigurd Svendsen Jeanette Hammer Andersen Johan Svenson 《Journal of molecular recognition : JMR》2013,26(10):461-469
Several drugs interact with the major plasma proteins serum albumin and alpha‐1 acid glycoprotein. Such binding may be either beneficial or disadvantageous from a pharmacokinetic perspective. In the present paper, we investigate the thermodynamics involved in the binding of a series of promising cationic antimicrobial peptides to the alpha‐1 acid glycoprotein using isothermal titration calorimetry. The drug‐like peptides are able to effectively destroy multiresistant bacterial strains, and members of this peptide class are currently in clinical phase II trials. Similar peptides, in a previous study, have been shown to bind to serum albumin resulting in a 10‐fold reduction in the peptides ability to kill bacteria in vitro. Here, it is shown that the peptides also are ligands for alpha‐1 glycoprotein with moderate binding affinities. The binding mode is investigated in detail using molecular docking, which maps the interaction to sub‐pockets I, II and III of the binding site. Despite this interaction, protein binding is shown to have little or no effect on the ability of the peptides to kill bacteria in vitro, either at normal physiological or acute phase concentrations. The results show that although the peptides interact with the binding pocket of alpha‐1 acid glycoprotein, the low stoichiometric binding ratio ensures that the interaction is not an obstacle for further development of these promising peptides as antimicrobial therapies. Copyright © 2013 John Wiley & Sons, Ltd. 相似文献
19.
Ansamitocins in combination with amphotericin B produced synergistic inhibition on the growth of several yeasts in liquid cultures, Ansamitocin P–3 at 5 µg/ml completely suppressed the growth of Saccharomyces cerevisiae whereas ansamitocin P–3 alone at 50 µg/ml hardly affected growth. Ansamitocin P–4 and maytansine also showed synergistic activity with amphotericin B against S. cerevisiae. The synergism also occurred in cultures of Candida albicans and Hansenula anomala. Combinations of ansamitocin P–3 with various agents revealed that the synergism depended on the specific property of amphotericin B. Ansamitocins showed no interfering activity against regeneration of protoplasts of S. cerevisiae. These results suggest that the limited activity of ansamitocins against these yeasts is due to the membrane permeability barrier of these cells. 相似文献
20.
Two aspects were studied to elucidate the functional and structural characterization of apidaecin and its N-terminal and C-terminal fragments: (i) Functions of the N-terminal and C-terminal fragments of apidaecin were first studied by measuring their antibacterial activity, their ability to enter Escherichia coli cells and their effects on the activities of beta-galactosidase and alkaline phosphatase. The results indicate that neither the N-terminal nor the C-terminal of apidaecin contains intracellular delivery unit or active segment. (ii) The effect of apidaecin on the ATPase activity of DnaK, and the interactions of apidaecin with E.coli lidless DnaK and DnaK D-E helix were studied. Results showed that apidaecin could interact with the E.coli lidless DnaK protein and stimulate its ATPase activity, but not with E.coli DnaK D-E helix. This indicated that the antimicrobial activity of apidaecin may be shown by stimulating the ATPase activity of DnaK by binding to its conventional substrate-binding site, to decrease its cellular concentration of DnaK by competing with natural substrates and inhibit the enzymes' activities of E. coli cells. It is the first study to suggest that the apidaecin-binding site of DnaK is the conventional substrate binging site. 相似文献