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 890 毫秒
1.
Conformations of ribostamycin and isepamicin, aminoglycoside antibiotics, bound to an aminoglycoside antibiotic, 3′-phosphotransferase, were determined by transferred nuclear Overhauser effect spectroscopy and molecular modeling. Two major conformers of enzyme-bound ribostamycin, a neomycin-group aminoglyeoside were observed. The 3′- and 5″-OH groups (reactive hydroxyl groups) in the conformers are placed in approximate locations. One of the conformers is similar to the structure of paromomycin bound to a 27-nucleotide piece of ribosomal RNA that represents the A-site of the small ribosomal subunit, where rings A and C are in an orthogonal arrangement. Isepamicin, a kanamycin-group aminoglycoside antibiotic, also showed two major enzyme-bound conformations. Both conformations were similar to those observed for bound isepamicin in the active site of an aminoglycoside(6′)-acetyl transferase-Ii. Conformations of other RNA-bound kanamycin-group aminoglycosides were also similar to the enzyme-bound conformations of isepamicin. These observations suggest that aminoglycosides may adopt similar conformations when bound to RNA and protein targets. This may have significant implications in the design of enzyme inhibitors and/or antibiotics.  相似文献   

2.
Aminoglycoside nucleotidyltransferase (2')-Ia [ANT (2')-Ia] was cloned from Pseudomonas aeruginosa and purified from overexpressing Escherichia coli BL21(DE3) cells. The first enzyme-bound conformation of an aminoglycoside antibiotic in the active site of an aminoglycoside nucleotidyltransferase was determined using the purified aminoglycoside nucleotidyltransferase (2' ')-Ia. The conformation of the aminoglycoside antibiotic isepamicin, a psuedo-trisaccharide, bound to aminoglycoside nucleotidyltransferase (2' ')-Ia has been determined using NMR spectroscopy. Molecular modeling, employing experimentally determined interproton distances, resulted in two different enzyme-bound conformations (conformer 1 and conformer 2) of isepamicin. Conformer 1 was by far the major conformer defined by the following average glycosidic dihedral angles: PhiBC = -65.26 +/- 1.63 degrees and PsiBC = -54.76 +/- 4.64 degrees. Conformer 1 was further subdivided into one major (conformer 1a) and two minor components (conformers 1b and 1c) based on the comparison of glycosidic dihedral angles PhiAB and PsiAB. The arrangement of substrates in the enzyme.metal-ATP.isepamicin complex was determined on the basis of the measured effect of the paramagnetic substrate analogue Cr(H2O)4ATP on the relaxation rates of substrate protons which were used to determine relative distances of isepamicin protons to the Cr3+. Both conformers of isepamicin yielded arrangements that satisfied the NOE restraints and the observed paramagnetic effects of Cr(H2O)4ATP. It has been suggested that aminoglycosides use both electrostatic interactions and hydrogen bonds in binding to RNA and that the contacts made by the A and B rings to RNA are the most important for binding [Fourmy, D., Recht, M. I., Blanchard, S. C., and Puglisi, J. D. (1996) Science 274, 1367-1371]. Comparisons based on the determined conformations of enzyme-bound aminoglycoside antibiotics also suggested that interactions of rings A and B with enzymes may be the major determinant in aminoglycoside binding to enzymes [Serpersu, E. H., Cox, J. R., DiGiammarino, E. L., Mohler, M. L., Ekman, D. R., Akal-Strader, A., and Owston, M. (2000) Cell Biochem. Biophys. (in press)]. The conformation of isepamicin bound to the aminoglycoside nucleotidyltransferase (2' ')-Ia, determined in this work, lent further support to this theory. Furthermore, comparison of enzyme-bound conformations of isepamicin to the RNA-bound conformation of gentamycin C1a also showed remarkable similarities between the enzyme-bound and RNA-bound aminoglycoside antibiotic conformations. These studies should aid in the design of effective inhibitors possessing a broad range of aminoglycoside-modifying enzymes as targets.  相似文献   

3.
Streptomycestenjimariensis SS-939 was resistant to its own aminoglycoside antibiotics, istamycins, as well as kanamycin A, neamine, ribostamycin and butirosin A, but was susceptible to neomycin B, lividomycin A and streptomycin. This resistance to these antibiotics was found to be due to ribosomes of the strain.  相似文献   

4.
The aminoglycoside phosphotransferases (APHs) are widely distributed among pathogenic bacteria and are employed to covalently modify, and thereby detoxify, the clinically relevant aminoglycoside antibiotics. The crystal structure for one of these aminoglycoside kinases, APH(3')-IIIa, has been determined in complex with ADP and analysis of the electrostatic surface potential indicates that there is a large anionic depression present adjacent to the terminal phosphate group of the nucleotide. This region also includes a conserved COOH-terminal alpha-helix that contains the COOH-terminal residue Phe(264). We report here mutagenesis and computer modeling studies aimed at examining the mode of aminoglycoside binding to APH(3')-IIIa. Specifically, seven site mutants were studied, five from the COOH-terminal helix (Asp(261), Glu(262), and Phe(264)), and two additional residues that line the wall of the anionic depression (Tyr(55) and Arg(211)). Using a molecular modeling approach, six ternary complexes of APH(3')-IIIa.ATP with the antibiotics, kanamycin, amikacin, butirosin, and ribostamycin were independently constructed and these agree well with the mutagenesis data. The results obtained show that the COOH-terminal carboxylate of Phe(264) is critical for proper function of the enzyme. Furthermore, these studies demonstrate that there exists multiple binding modes for the aminoglycosides, which provides a molecular basis for the broad substrate- and regiospecificity observed for this enzyme.  相似文献   

5.
L J Ferrin  A S Mildvan 《Biochemistry》1986,25(18):5131-5145
The large fragment of DNA polymerase I (Pol I) effectively uses oligoribouridylates and oligoriboadenylates as templates, with kinetic properties similar to those of poly(U) and poly(A), respectively, and has little or no activity in degrading them. In the presence of such oligoribonucleotide templates, nuclear Overhauser effects (NOE's) were used to determine interproton distances within and conformations of substrates bound to the large fragment of Pol I, as well as conformations and interactions of the enzyme-bound templates. In the enzyme-oligo(rU)54 +/- 11-Mg2+dATP complex, the substrate dATP has a high anti-glycosidic torsional angle (chi = 62 +/- 10 degrees) and an O1'-endo/C3'-endo sugar pucker (delta = 90 +/- 10 degrees) differing only slightly from those previously found for enzyme-bound dATP in the absence of template [Ferrin, L.J., & Mildvan, A.S. (1985) Biochemistry 24, 4680-4694]. Both conformations are similar to those of deoxynucleotidyl units of B DNA but differ greatly from those of A or Z DNA. The conformation of the enzyme-bound substrate analogue AMPCPP (chi = 50 +/- 10 degrees, delta = 90 +/- 10 degrees) is very similar to that of enzyme-bound dATP and is unaltered by the binding of the template oligo(rU)54 +/- 11 or by the subsequent binding of the primer (Ap)9A. In the enzyme-oligo(rA)50-Mg2+TTP complex, the substrate TTP has an anti-glycosidic torsional angle (chi = 40 +/- 10 degrees) and an O1'-endo sugar pucker (delta = 100 +/- 10 degrees), indistinguishable from those found in the absence of template and compatible with those of B DNA but not with those of A or Z DNA. In the absence of templates, the interproton distances on enzyme-bound dGTP cannot be fit by a single conformation but require a 40% contribution from a syn structure (chi = 222 degrees) and a 60% contribution from one or more anti structures. The presence of the template oligo(rU)43 +/- 9 simplifies the conformation of enzyme-bound dGTP to a single structure with an anti-glycosyl angle (chi = 32 +/- 10 degrees) and an O1'-endo/C3'-endo sugar pucker (delta = 90 +/- 10 degrees), compatible with those of B DNA, possibly due to the formation of a G-U wobble base pair. However, no significant misincorporation of guanine deoxynucleotides by the enzyme is detected with oligo(rU) as template.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

6.
AAC(2')-Ic catalyzes the coenzyme A (CoA)-dependent acetylation of the 2' hydroxyl or amino group of a broad spectrum of aminoglycosides. The crystal structure of the AAC(2')-Ic from Mycobacterium tuberculosis has been determined in the apo enzyme form and in ternary complexes with CoA and either tobramycin, kanamycin A or ribostamycin, representing the first structures of an aminoglycoside acetyltransferase bound to a drug. The overall fold of AAC(2')-Ic places it in the GCN5-related N-acetyltransferase (GNAT) superfamily. Although the physiological function of AAC(2')-Ic is uncertain, a structural analysis of these high-affinity aminoglycoside complexes suggests that the enzyme may acetylate a key biosynthetic intermediate of mycothiol, the major reducing agent in mycobacteria, and participate in the regulation of cellular redox potential.  相似文献   

7.
The angiotensin I-converting enzyme (peptidyl-dipeptide hydrolase, EC 3.4.15.1) inhibitor, ramiprilat (2-[N-[(S)-1-ethoxycarbonyl-3-phenylpropyl]-L-Ala]-(1S,3S,5S)-2- azabicyclo[3.3.0]octane-3-carboxylic acid), is shown to exist in tow conformational isomers, cis and trans, which interconvert around the amide bond. The two conformers were separated by reversed-phase high-performance liquid chromatography. The conformers were identified by nuclear Overhauser effect measurements. From line shape analysis the isomerization rate constants were determined to be kcis----trans = 15 s-1 and ktrans----cis = 5 s-1 at 368 K in [2H]phosphate buffer (p2H 7.5). By enzyme kinetic studies using 3-(2-furylacryloyl)-L-Phe-Gly-Gly as substrate, the trans conformer was found to be the most potent enzyme inhibitor, whereas the cis conformer had a very low inhibitory effect. A new inhibition mechanism is presented for this type of slow, tight-binding inhibitors that contain an amide bond. This mechanism involves an equilibrium between the two conformers and the enzyme-bound inhibitor complex.  相似文献   

8.
The crystal structures of six complexes between aminoglycoside antibiotics (neamine, gentamicin C1A, kanamycin A, ribostamycin, lividomycin A and neomycin B) and oligonucleotides containing the decoding A site of bacterial ribosomes are reported at resolutions between 2.2 and 3.0 Å. Although the number of contacts between the RNA and the aminoglycosides varies between 20 and 31, up to eight direct hydrogen bonds between rings I and II of the neamine moiety are conserved in the observed complexes. The puckered sugar ring I is inserted into the A site helix by stacking against G1491 and forms a pseudo base pair with two H-bonds to the Watson–Crick sites of the universally conserved A1408. This central interaction helps to maintain A1492 and A1493 in a bulged-out conformation. All these structures of the minimal A site RNA complexed to various aminoglycosides display crystal packings with intermolecular contacts between the bulging A1492 and A1493 and the shallow/minor groove of Watson–Crick pairs in a neighbouring helix. In one crystal, one empty A site is observed. In two crystals, two aminoglycosides are bound to the same A site with one bound specifically and the other bound in various ways in the deep/major groove at the edge of the A sites.  相似文献   

9.
Aminoglycoside antibiotics are frequently used for the selection of transgenic plant cells. However, for a number of species aminoglycoside selection is inefficient. The objective of the present study was to elucidate factors affecting the phytoloxic effects of aminoglycoside antibiotics. Using non-transgenic sugar beet cotyledonary explants the interaction between three aminoglycoside antibiotics, kanamycin, neomycin and hygromycin. and Ca2+ was studied by monitoring the effects on growth and shoot formation. The phytotoxic effects of the aminoglycoside antibiotics were strongly dependent on the calcium concentration in the growth media. At comparable levels of the antibiotics (kanamycin 170 μ M , neomycin 220 μ M , hygromycin 9.5 μ M , an elevation of the calcium concentration from 1 to 10 m M resulted in growth increases of approximately 3-, 2.5- and 8-fold, respectively, and shoot formation was enhanced 1.5-, 2-and 6-fold, respectively. At lower concentrations of the antibiotics, the toxic effect was nearly abolished by increasing the calcium concentration. Additional magnesium, sodium and ammonium did not affect the phytotoxic effects of the aminoglycoside antibiotics. Moreover, the phytotoxic effects of the herbicides glyphosate and phosphinothricin were not decreased by additional calcium. These data suggest the existence of a specific interaction between calcium and aminoglycoside anfibiotics in plants. The implications of these results for the use of aminoglycosides as selective agents in plant transformation are discussed.  相似文献   

10.
A transposable element, Tn2401, was found in a clinical isolate of Pseudomonas aeruginosa. Tn2401 had a size of 7190 nucleotides and encoded aminoglycoside 3'-phosphotransferase and aminoglycoside 6'-N-acetyltransferase. The sequence encoding the former enzyme was homologous with that of Tn903. Pseudomonas aeruginosa strains harbouring this transposon were resistant to kanamycin, neomycin, lividomycin, ribostamycin, paromomycin, netilmycin, tobramycin, dibekacin, gentamicin, sisomicin, and butirosin.  相似文献   

11.
Endogenous lectins induce effects on cell growth by binding to antennae of natural glycoconjugates. These complex carbohydrates often present more than one potential lectin-binding site in a single chain. Using the growth-regulatory interaction of the pentasaccharide of ganglioside GM(1) with homodimeric galectin-1 on neuroblastoma cell surfaces as a model, we present a suitable strategy for addressing this issue. The approach combines NMR spectroscopic and computational methods and does not require isotope-labeled glycans. It involves conformational analysis of the two building blocks of the GM(1) glycan, i.e., the disaccharide Galbeta1-3GalNAc and the trisaccharide Neu5Acalpha2-3Galbeta1-4Glc. Their bound-state conformations were determined by transferred nuclear Overhauser enhancement spectroscopy. Next, measurements on the lectin-pentasaccharide complex revealed differential conformer selection regarding the sialylgalactose linkage in the tri- versus pentasaccharide (Phi and Psi value of -70 degrees and 15 degrees vs 70 degrees and 15 degrees, respectively). To proceed in the structural analysis, the characteristic experimentally detected spatial vicinity of a galactose unit and Trp68 in the galectin's binding site offered a means, exploiting saturation transfer from protein to carbohydrate protons. Indeed, we detected two signals unambiguously assigned to the terminal Gal and the GalNAc residues. Computational docking and interaction energy analyses of the entire set of ligands supported and added to experimental results. The finding that the ganglioside's carbohydrate chain is subject to differential conformer selection at the sialylgalactose linkage by galectin-1 and GM(1)-binding cholera toxin (Phi and Psi values of -172 degrees and -26 degrees, respectively) is relevant for toxin-directed drug design. In principle, our methodology can be applied in studies aimed at blocking galectin functionality in malignancy and beyond glycosciences.  相似文献   

12.
AIMS: To investigate the in vitro antifungal and antioomycete activities of some aminoglycosides against true fungi and Phytophthora and Pythium species and to evaluate the potential of the antibiotics against Phytophthora late blight on plants. METHODS AND RESULTS: Antifungal and antioomycete activities of aminoglycoside antibiotics (neomycin, paromomycin, ribostamycin and streptomycin) and a paromomycin-producing strain (Streptomyces sp. AMG-P1) against Phytophthora and Pythium species and 10 common fungi were measured in potato dextrose broth (PDB) and on seedlings in pots. Paromomycin was the most active against Phytophthora and Pythium species with a minimal inhibitory concentration of 1-10 microg ml(-1) in PDB, but displayed low to moderate activities towards other common fungi at the same concentration. Paromomycin also showed potent in vivo activity against red pepper and tomato late blight diseases with 80 and 99% control value, respectively, at 100 microg ml(-1). In addition, culture broth of Streptomyces sp. AMG-P1 as a paromomycin producer exhibited high in vivo activity against late blight at 500 microg freeze-dried weight per millilitre. CONCLUSIONS: Among tested aminoglycoside antibiotics, paromomycin was the most active against oomycetes both in vitro and in vivo. SIGNIFICANCE AND IMPACT OF THE STUDY: Data from this study show that aminoglycoside antibiotics have in vitro and in vivo activities against oomycetes, suggesting that Streptomyces sp. AMG-P1 may be used as a biocontrol agent against oomycete diseases.  相似文献   

13.
Gad GF  Mohamed HA  Ashour HM 《PloS one》2011,6(2):e17224
With the re-emergence of older antibiotics as valuable choices for treatment of serious infections, we studied the aminoglycoside resistance of Gram-negative bacteria isolated from patients with ear, urinary tract, skin, and gastrointestinal tract infections at Minia university hospital in Egypt. Escherichia coli (mainly from urinary tract and gastrointestinal tract infections) was the most prevalent isolate (28.57%), followed by Pseudomonas aeruginosa (25.7%) (mainly from ear discharge and skin infections). Isolates exhibited maximal resistance against streptomycin (83.4%), and minimal resistance against amikacin (17.7%) and intermediate degrees of resistance against neomycin, kanamycin, gentamicin, and tobramycin. Resistance to older aminoglycosides was higher than newer aminoglycosides. The most common aminoglycoside resistance phenotype was that of streptomycin resistance, present as a single phenotype or in combination, followed by kanamycin-neomycin as determined by interpretative reading. The resistant Pseudomonas aeruginosa strains were capable of producing aminoglycoside-modifying enzymes and using efflux as mechanisms of resistance. Using checkerboard titration method, the most frequently-observed outcome in combinations of aminoglycosides with β-lactams or quinolones was synergism. The most effective combination was amikacin with ciprofloxacin (100% Synergism), whereas the least effective combination was gentamicin with amoxicillin (53.3% Synergistic, 26.7% additive, and 20% indifferent FIC indices). Whereas the studied combinations were additive and indifferent against few of the tested strains, antagonism was never observed. The high resistance rates to aminoglycosides exhibited by Gram-negative bacteria in this study could be attributed to the selective pressure of aminoglycoside usage which could be controlled by successful implementation of infection control measures.  相似文献   

14.
The major mechanism of resistance to aminoglycosides in clinical bacterial isolates is the covalent modification of these antibiotics by enzymes produced by the bacteria. Aminoglycoside 2'-Ib phosphotransferase [APH(2')-Ib] produces resistance to several clinically important aminoglycosides in both Gram-positive and Gram-negative bacteria. Nuclear magnetic resonance analysis of the product of kanamycin A phosphorylation revealed that modification occurs at the 2'-hydroxyl of the aminoglycoside. APH(2')-Ib phosphorylates 4,6-disubstituted aminoglycosides with kcat/Km values of 10(5)-10(7) M-1 s-1, while 4,5-disubstituted antibiotics are not substrates for the enzyme. Initial velocity studies demonstrate that APH(2')-Ib operates by a sequential mechanism. Product and dead-end inhibition patterns indicate that binding of aminoglycoside antibiotic and ATP occurs in a random manner. These data, together with the results of solvent isotope and viscosity effect studies, demonstrate that APH(2')-Ib follows the random Bi-Bi kinetic mechanism and substrate binding and/or product release could limit the rate of reaction.  相似文献   

15.
The present paper describes the predominant conformational forms adopted by dipeptides in aqueous solution. More than 50 dipeptides were subjected to conformational analysis using SYBYL Random Search. The resultant collections of conformers for individual dipeptides, for small groups with related side chain residues and for large groups of about 50 dipeptides were visualized graphically and analysed using a novel three-dimensional pseudo-Ramachandran plot. The distribution of conformers, weighted according to the percentage of each in the total conformer pool, was found to be restricted to nine main combinations of backbone psi (psi) and phi (phi) torsion angles. The preferred psi values were in sectors A7 (+150 degrees to +/-180 degrees), A10 (+60 degrees to +90 degrees) and A4 (-60 degrees to -90 degrees), and these were combined with preferred phi values in sectors B12 (-150 degrees to +/-180 degrees), B9 (-60 degrees to -90 degrees) and B2 (+30 degrees to +60 degrees). These combinations of psi and phi values are distinct from those found in common secondary structures of proteins. These results show that although dipeptides can each adopt many conformations in solution, each possesses a profile of common conformers that is quantifiable. A similarly weighted distribution of dipeptide conformers according to distance between amino-terminal nitrogen and carboxyl-terminal carbon shows how the preferred combinations of backbone torsional angles result in particular N-C geometries for the conformers. This approach gives insight into the important conformational parameters of dipeptides that provide the basis for their molecular recognition as substrates by widely distributed peptide transporters. It offers a basis for the rational design of peptide-based bioactive compounds able to exploit these transporters for targeting and delivery.  相似文献   

16.
17.
The FTIR spectra were analysed in the region of the nu(s)(N-H), AI(C=O) and nu(s)(Calpha=Cbeta) bands for a series of Ac-DeltaXaa-NMe2, where DeltaXaa = DeltaAla, (Z)-DeltaAbu, (Z)-DeltaLeu, (Z)-DeltaPhe and DeltaVal, to determine a predominant solution conformation of these alpha,beta-dehydropeptide-related molecules. Measurements were taken in CCl4, DCM and MeCN solutions. In the same way, spectra of saturated analogues Ac-Xaa-NMe2, where Xaa = Ala, Abu, Leu, Phe and Val, were investigated. To help interpret the spectroscopic results, conformational maps were calculated by the B3LYP/6-31+G** method. Also, the relative energies of all conformers of the dehydro compounds in vacuo as well as in the studied solvents in addition to the theoretical IR frequencies of these conformers were calculated. For comparison, molecules of two saturated analogues, Ac-L-Ala-NMe2 and Ac-L-Phe-NMe2, were calculated in a similar way. Both unsaturated and saturated compounds, which have an aliphatic side chain, occur in CCl4 and DCM mainly as a mixture of extended conformers with the C5 H-bond and open conformers. As solvent polarity increases, participation of the open conformers also increases, and in MeCN, the model amides are almost exclusively in the open form, except Ac-DeltaAla-NMe2, which shows a small amount of the H-bonded conformer. Ac-DeltaAla-NMe2 and Ac-DeltaAbu-NMe2 have stronger C5 hydrogen bonds than those of their saturated counterparts. As the calculations indicate, the open conformation of the unsaturated amides is conformer H/F with phi, psi -44 +/- 5 degrees, 127 +/- 4 degrees. This is the second lowest in energy conformer in vacuo and in CCl4 and the lowest one in more polar solvents. The open conformation of Ac-L-Ala-NMe2 constitutes conformer C with phi, psi -101.5 degrees, 112.7 degrees. For Ac-DeltaAla-NMe2 and Ac-DeltaAbu-NMe2, FTIR also reveals the presence of a third conformer. Calculations indicate that is the semiextended conformer D with the N1-H1...N2 hydrogen bond/contact. In all solvents, Ac-L-Phe-NMe2 and Ac-(Z)-DeltaPhe-NMe2 show only the extended E and the open H/F, respectively. In both there is an amide/pi(Ph) interaction.  相似文献   

18.
The binding of methyl beta-lactoside and of all possible monodeoxy derivatives of methyl beta-lactoside to the galactose-specific highly cytotoxin lectin ricin, has been investigated. The distribution of low-energy conformers of the disaccharide structures has been first determined using molecular-mechanics calculations and high-resolution NMR spectroscopy. The nuclear Overhauser enhancements and specific deshieldings observed are in agreement with a similar distribution of low-energy conformers for all studied compounds which may be described by a major conformer defined by phi (H1'-C1'-O1'-C4) and psi (C1'-O1'-C4-H4) torsion angles of 49 degrees and 5 degrees, respectively, with contribution of conformers with angles phi/psi 24 degrees/-59 degrees, 22 degrees/-32 degrees and 6 degrees/-44 degrees. Assuming that the disaccharides bind to the lectin in these preferred conformations, the apparent dissociation constants for the ricin-disaccharide complexes have been interpreted in terms of specific polar and nonpolar interactions. In agreement with X-ray data, the hydroxyl groups at positions 3, 4 and 6 of the beta-D-galactopyranose moiety appear as key polar groups in the interaction with ricin. These results are in contrast to previous results which have established that position 6 is not involved in lectin binding. An important nonpolar interaction involving position 3 of the beta-D-glucopyranose moiety, seems to be operative. The distribution of low-energy conformers of these disaccharide structures permits this interaction to take place with the hydroxyl group at this position intramolecularly bonded, thus rendering this region of the molecule more lipophylic in character for acceptance into nonpolar regions of the combining site.  相似文献   

19.
Kaul M  Pilch DS 《Biochemistry》2002,41(24):7695-7706
We use spectroscopic and calorimetric techniques to characterize the binding of the aminoglycoside antibiotics neomycin, paromomycin, and ribostamycin to a RNA oligonucleotide that models the A-site of Escherichia coli 16S rRNA. Our results reveal the following significant features: (i) Aminoglycoside binding enhances the thermal stability of the A-site RNA duplex, with the extent of this thermal enhancement decreasing with increasing pH and/or Na(+) concentration. (ii) The RNA binding enthalpies of the aminoglycosides become more exothermic (favorable) with increasing pH, an observation consistent with binding-linked protonation of one or more drug amino groups. (iii) Isothermal titration calorimetry (ITC) studies conducted as a function of buffer reveal that aminoglycoside binding to the host RNA is linked to the uptake of protons, with the number of linked protons being dependent on pH. Specifically, increasing the pH results in a corresponding increase in the number of linked protons. (iv) ITC studies conducted at 25 and 37 degrees C reveal that aminoglycoside-RNA complexation is associated with a negative heat capacity change (Delta C(p)), the magnitude of which becomes greater with increasing pH. (v) The observed RNA binding affinities of the aminoglycosides decrease with increasing pH and/or Na(+) concentration. In addition, the thermodynamic forces underlying these RNA binding affinities also change as a function of pH. Specifically, with increasing pH, the enthalpic contribution to the observed RNA binding affinity increases, while the corresponding entropic contribution to binding decreases. (vi) The affinities of the aminoglycosides for the host RNA follow the hierarchy neomycin > paromomycin > ribostamycin. The enhanced affinity of neomycin relative to either paromomycin or ribostamycin is primarily, if not entirely, enthalpic in origin. (vii) The salt dependencies of the RNA binding affinities of neomycin and paromomycin are consistent with at least three drug NH(3)(+) groups participating in electrostatic interactions with the host RNA. In the aggregate, our results reveal the impact of specific alterations in aminoglycoside structure on the thermodynamics of binding to an A-site model RNA oligonucleotide. Such systematic comparative studies are critical first steps toward establishing the thermodynamic database required for enhancing our understanding of the molecular forces that dictate and control aminoglycoside recognition of RNA.  相似文献   

20.
Based on the structural information of biomacromolecule-aminoglycoside complexes, a series of kanamycin B analogues were rationally designed and synthesized. A convenient approach to the construction of kanamycin derivatives, in which the C4′-position on ring I of neamine moiety was modified, was developed. Most synthetic analogues exhibited good to excellent antibiotic activity against some typical drug-resistant bacteria. The disclosed results suggested that the C4′-position of aminoglycosides such as kanamycin may be an ideal site for modification to gain new modifying enzyme-resistant aminoglycoside antibiotics.  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