GET THE APP

Effects of Proton Linkage on Thermodynamic Properties of Enzyme-A
Journal of Thermodynamics & Catalysis

Journal of Thermodynamics & Catalysis
Open Access

ISSN: 2157-7544

+44 1300 500008

Research Article - (2011) Volume 2, Issue 1

Effects of Proton Linkage on Thermodynamic Properties of Enzyme-Antibiotic Complexes of the Aminoglycoside Nucleotidyltransferase (2³)-Ia

Edward Wright1 and Engin H. Serpersu2*
1Department of Biochemistry and Cellular and Molecular Biology, The University of Tennessee-Knoxville, Knoxville, Tennessee 37996, USA
2University of Tennessee-Knoxville, Department of Biochemistry, Cellular and Molecular Biology, Walters Life Sciences Bldg. M407, Knoxville, TN 37996-0840, USA
*Corresponding Author: Dr. Engin H. Serpersu, University of Tennessee-Knoxville, Department of Biochemistry, Cellular and Molecular Biology, Walters Life Sciences Bldg, M407, Knoxville, TN 37996-0840, USA, Tel: 865-974-2668, Fax: 865-974- 6306 Email:

Abstract

Aminoglycoside nucleotidyltransferase(2″)-Ia (ANT) catalyzes the covalent modification of certain aminoglycoside antibiotics and imparts resistance to bacteria which possess this enzyme. Isothermal titration calorimetry (ITC) experiments using two structurally similar substrates (kanamycin A and kanamycin B) and a structurally different inhibitor (neomycin) were performed to determine pH and proton linkage effects on thermodynamic properties of enzyme–aminoglycoside complexes. Data showed that there was a decrease in pKa for one or more carboxyl groups on the enzyme that interacts with 2′-amino group of neomycin or kanamycin B. Protonation state of this site exerted a significant effect on the binding affinity of ligand to ANT. ITC experiments also showed that the loss of affinity at higher pH values was not as pronounced for neomycin compared to kanamycins. This difference is a result of higher pKas for free neomycin which are up shifted in the ANT–neomycin complex. The decrease in affinity and differences in binding linked protonation above and below the pH optimum cannot be fully explained by these two factors. Therefore, the formation of ANT–aminoglycoside complexes involves protonation and deprotonation of multiple functional groups on both ligand and enzyme which also contribute to the pH profile of the complex formation.

Introduction

Aminoglycosides are broad-spectrum antibiotics that bind to bacterial 16S rRNA. This binding causes an increase in proof-reading errors and mistranslation which eventually leads to disruption of the cell membrane and death of the microorganism. Structurally, these antibiotics consist of a 2-deoxystreptamine ring with amino sugar substitutions at either the 4,5 or 4,6 positions (Figure 1). Streptomycin is an exception to the common aminoglycoside chemical motif. These antibiotics are used to treat serious infections caused by Gram-negative bacteria.

Enzymatic modification of aminoglycoside antibiotics is the major resistance mechanism to the action of these antibiotics [1]. The aminoglycoside nucleotidyltransferase(2″)-Ia (ANT) is one of the most clinically important aminoglycoside modifying enzymes (AGMEs). This enzyme catalyzes the adenylation of aminoglycosides which contain a hydroxyl group at the 2″ position. ANT is one of only four of the over fifty known AGMEs to modify primarily at positions on ring C of aminoglycosides (Figure 1) [2]. Most AGMEs modify positions on the A or B ring.

thermodynamics-catalysis-Alternate-letter-labeling

Figure 1: Aminoglycosides Kanamycins A and B (top) and Neomycin (bottom). Substitution at the 2′-site for kanamycin B is shown in parenthesis (top). Amino groups are shown in blue with pKas indicated below. Numbering are shown in parenthesis on the side. Alternate letter-labeling of the rings are also shown.

ANT was first reported in a clinical isolate of Klebsiella pneumoniae in 1971 [3]. By 1985 ANT was the most often detected aminoglycoside modifying enzyme in resistant clinical isolates in North America and had been found in each of the major continents [4,5]. ANT continues to be one of the two most prevalent enzymatic aminoglycoside resistance mechanisms in Pseudomonas aeruginosa [6,7]. Despite its medical importance, very little is known about this enzyme. This protein is difficult to isolate in large quantities with a high level of purity [8,9]. It is also relatively unstable and has a low degree of solubility [10]. Because of these problems, earlier studies focused on kinetic assays and in vivo inhibition patterns of this enzyme [8,11-13]. More recently, analysis of thermodynamic properties of ANT–aminoglycoside complexes showed that all binding interactions were characterized by favorable enthalpy (−ΔH) and unfavorable entropy (−TΔS) [14,15], which is typical for carbohydrate-protein interactions [16-18]. This study also looked at substrate selectivity by ANT [15].

To further investigate the thermodynamics of ANT–aminoglycoside interactions the role of proton linkage and pH dependence of substrate binding was investigated. Two different calorimetric approaches were used to determine binding-linked protonation/deprotonation and the effect of pH on the association constant (Kb) of the binary ANT− aminoglycoside complexes. In the first method, enthalpy change (ΔHobs) upon ligand (aminoglycoside) binding to the enzyme in buffers with different ionization enthalpies (ΔHion) was determined to derive intrinsic enthalpy change (ΔHint) and the net number of bindingcoupled proton exchange (Δn). When pKa shifts occur in either the protein or ligand, uptake or release of protons to the buffer will occur. This exchange will contribute to the observed ΔH and must be considered in the thermodynamic analysis of the system. The binding of aminoglycosides to AGMEs include significant contribution to the observed enthalpy of complex formation from the protonation and deprotonation of titratable groups on the enzyme and ligands [15,19,20]. In the second method, pH dependence of association was measured using a single buffer system. This buffer system encompassed the pH range of Kb values accurately measurable by ITC for ANT– aminoglycoside interactions. This paper shows that multiple groups on ligand and enzyme shift their pKas upon formation of the binary ANT–aminoglycoside complexes.

Materials and Methods

Reagents

All materials were of the highest purity commercially available. All were purchased from Sigma-Aldrich Co. (St. Louis, MO) except for tris(2-carboxyethyl)phosphine (TCEP) purchased from Fluka (Buchs, Switzerland). ANT was prepared as described previously [10]. Activity assays of the enzyme were also performed as described previously [14].

Isothermal titration calorimetry

ITC experiments were performed at 20°C using a VP-ITC microcalorimeter from Microcal, Inc. (Northampton, MA).The enzyme was exchanged into the appropriate buffer using a Sephadex G-25 column and the ligand solution was prepared using the same buffer solution. The final buffer for determination of pH dependence was 50 mM MES/Tris, 50 mM KCl and 2 mM TCEP. The buffers for determination of the net uptake of protons contained 50 mM of the indicated buffer. The following buffers with the enthalpy of ionization in kcal mol-1 in parenthesis were used at pH 7.1: PIPES (2.68), HEPES (4.87), MOPS (5.04) and ACES (7.26), at pH 7.6: HEPES (4.87), MOPS (5.04), Bicine (6.29) and Tris (11.3), and at pH 8.1: HEPES (4.87), Bicine (6.29), TAPS (9.65) and Tris (11.3) [21]. The concentration of aminoglycoside was determined by NMR and activity assays as described previously [14]. Both enzyme and ligand solutions were degassed under vacuum for 10 min at 15°C. Titrations consisted of 29 injections of 10 μL and were separated by 240 s. Cell stirring speed was 300 rpm. Each titration contained 5–40 μM enzyme in the sample cell. The aminoglycoside concentration was 60–800 μM in the injection syringe. The standard errors represent the deviation including curve fitting errors of the three titrations. The c values (c = KbMt, where Mt is the concentration of macromolecule binding sites) were in the range 10–30 for experiments in the pH range 6.5–7.8. This range is ideal for accurately determining binding constants by ITC [22]. For titrations above pH 7.8 and below pH 6.5 the c values were in the range 0.4–10. The use of less than optimum c values was necessitated by the low affinity at these pH values combined with the limited solubility and strong tendency of oligomerization of this enzyme above 40 μM (~1 mg/ml) concentration. Monomeric state of the protein was confirmed by analytical size-exclusion chromatography. For all complexes observed in this study the thermodynamic parameters determined were independent of protein concentration.

The heat of dilution for each aminoglycoside was determined by titrating aminoglycoside into buffer in the absence of enzyme and subtracted prior to curve fitting. For all titrations the pH was confirmed immediately prior to the start of the experiment. All data were fit to the single-site binding model of Origin 5.0 (Microcal, Inc) to determine the binding constant (Kb), enthalpy of binding (ΔH) and stoichiometry (n). The free energy (ΔG) and entropy (ΔS) changes associated with binding were determined using the equations:

          ΔG = −RT Kb

         ΔG = ΔH −TΔS

Results and Discussion

ANT is one of the “less promiscuous” AGMEs and modifies only kanamycins and gentamicins while neomycins are competitive inhibitors of this enzyme. In this study, three different aminoglycoside antibiotics representing kanamycins and neomycins were used to determine thermodynamic properties of their complexes with ANT as a function of pH. Kanamycin A and kanamycin B were selected for this study because the thermodynamic parameters of their complexes with ANT show significant differences despite a minimal structural difference between them (-NH2 versus –OH at the 2′-site in kanamycin B and kanamycin A respectively) (Figure 1) [15]. Neomycin B was chosen as the representative of the neomycin class because of the high affinity of this antibiotic for ANT [15].

Effect of complex formation on the ionization of functional groups: In the presence of binding-linked protonation, the observed enthalpy (ΔHobs) includes contribution from various sources according to the equation [23]:

          ΔHobs = ΔHint + Δn [α ΔHion + (1-α)ΔHenz ] + ΔHbind

in which ΔHint is the intrinsic enthalpy of binding and Δn represents the net proton transfer. ΔHobs denotes the observed binding enthalpy of complex formation in a buffer where ΔHion describes the heat of ionization of the buffer. The term Δn [α ΔHion + (1-α)ΔHenz ] represents the heat of ionization of groups from the ionization of buffer and the protein to maintain pH, where α represent fraction of protonation contributed by the buffer [23]. In addition, ΔHbind represents the heat of binding of buffer to the enzyme. In the presence of sufficient salt (i.e., 50 mM KCl), ΔHbind is assumed to be zero and the contribution from the ionization of amino acids remains the same at a given pH. Thus, by performing experiments in buffers with different heats of ionization, one can easily determine ΔHint and Δn, which is based on the simplified version of the above equation (ΔHobs = ΔHint + Δn ΔHion). An example is shown in Figure 2 where the heat of ionization of buffers is plotted against ΔHobs for the formation of the ANT–kanamycin B complex at pH 7.1 and 7.6. The values for Δn and were determined from the slope and the intercept respectively. However, note that ΔHint still includes the heat of ionization of groups contributing to Δn (i.e., ΔHint = (ΔHint + ΔHfunctional groups · Δn) which would represent the true ΔHint only when Δn=0). For the buffers used in this work, a net proton uptake by enzyme–ligand complex yields a positive Δn value. Bindinglinked proton uptake was observed in the formation of the binary ANT–aminoglycoside complexes with the exception of one condition; ANT–neomycin complex yielded a negative Δn at pH 7.1 indicating a net release of protons from the complex under these conditions (i.e., lowered pKas) (Table 1). Values of ΔHintand Δn showed strong dependence on pH as shown in Table 1 and Figure 3. Observed positive values for Δn (Δn > 0) for the enzyme-aminoglycoside interactions suggests that several functional groups may have up shifted pKas in the binary ANT–aminoglycoside complexes. It is also reasonable to assume that increased pKa of amino groups in ligands contribute significantly to the observed positive value of Δn. Consistent with this, pKas of several amino groups in enzyme-bound neomycin showed up to one pKa unit higher shifts when bound to another aminoglycosidemodifying enzyme [24].

thermodynamics-catalysis-binary-enzyme

Figure 2: Change in the observed enthalpy of the binary enzyme–kanamycin B complex at pH 7.1 (○) and at pH 7.6 (■) as a function of the heat of ionization of different buffers.

  pH Δn ΔHint (kcal/mol)
Kan B 7.1 0.16 ± 0.12 -17.3 ± 0.9
Kan B 7.6 1.3 ± 0.3 -25.6 ± 1.7
Kan B 8.1 1.1 ± 0.2 -25.8 ± 1.2
Kan A 7.1 0.45  ± 0.15 -19.7 ± 1.8
Kan A 7.6 1.4 ± 0.2 -24.4 ± 0.6
Kan A 8.1  0.90  ± 0.2 -24.3 ± 2.8
Neomycin 7.1 -0.22  ± 0.14 -10.0  ± 1.3
Neomycin 7.6 1.0  ± 0.3 -18.9 ± 1.9
Neomycin 8.1 2.0 ± 0.2 -31.5 ± 1.1

Table 1: pH-dependent protonation in the binary ANT-Aminoglycoside complexes

ANT-aminoglycoside association as a function of pH

The second calorimetric method involves determination of the association constant (Kb) for the formation of binary enzyme– aminoglycoside complexes as a function of pH. A typical titration data demonstrating the effect of pH on the binding affinity of neomycin to ANT is shown in Figure 3 (please note the differences in y-axis scale). Effect of pH on the binding constant (Kb) of ANT–aminoglycoside complexes was determined by performing binding experiments in the pH range of 6.3 to 8.1. In general, all three aminoglycosides showed an increased affinity with increasing pH up to pH ~6.8. A plateau was observed at neutral pH region and the affinity started to decline above pH 7.5. Data for the binding of kanamycins and neomycin are shown in Figures 4, and 5 respectively.

thermodynamics-catalysis-fitted-data

Figure 3: Titration of ANT with neomycin at pH 7.4 (left panel) and pH 8.2 (right panel). Thermograms (upper panels) are shown with the fitted data (lower panels).

Binding of kanamycin A vs. kanamycin B to ANT

Kanamycin A and B are 4.6-disubstituted aminoglycosides which are identical except for the substituent at the 2′ position. Kanamycin B contains a 2′-NH2 while kanamycin A possesses a 2′-OH. The presence of an amino instead of a hydroxyl at this position results in a fivefold increase in affinity for kanamycin B relative to kanamycin A [15]. Binding of kanamycin A and B to ANT leads to a similar binding linked protonation profile. In both cases Δn increases from 7.1 to 7.6 then decreases from 7.6 to 8.1 (Table 1, Figure 4). At pH 7.6 and 8.1 the difference in Δn for the two complexes is within error suggesting that the 2′ position does not significantly affect Δn at these pH values. At pH 7.1 kanamycin B has a Δn value that is 0.3 less than the value for kanamycin A. If the binding interactions and energetic contributions remain consistent at the other positions in the two complexes, the difference must be attributable to the 2′-NH2 vs. 2′-OH. Crystal structures of aminoglycosides bound to AGMEs show the predominant interactions involve the amino groups on the substrate and aspartyl and to a lesser extent glutamyl side chains on the enzyme [25-27]. In structures of AGMEs which show a strong preference for kanamycin B relative to kanamycin A (similar to ANT) the 2′ position interacts with an acidic residue [25,27,28]. The most plausible explanation is that the contributions from the other positions on kanamycin result in a larger positive Δn for complex formation with ANT. The active site aspartate or glutamate that interacts with the 2′ position may have an up shifted pKa due to multiple acidic residues in the active site. The binding of kanamycin B with a positively charged amino group at the 2′ position will result in a larger decrease in pKa for the carboxyl group on the enzyme due to the favorability of a charge-charge interaction [29,30]. This effect would be magnified because the binding of aminoglycoside would be removing the carboxyl group from solvent [29]. If the hydroxyl group of kanamycin A forms a hydrogen bond with the same residue, it would also decrease the pKa of the side chain carboxyl, but to a lesser extent [29,31]. Although less likely, the difference in Δn at pH 7.1 could be due to a decrease in pKa of the 2′-NH2 of kanamycin B upon binding to ANT and an increase in affinity of 2′-NH2 compared to 2′-NH3 + for the enzyme, which also fits the data. In most cases, though, protonated amino groups lead to higher affinity for AGMEs [32]. Additionally, increases in pKa values for amino groups that are directly involved in binding to AGMEs have been observed previously [21].

thermodynamics-catalysis-overall-charge

Figure 4: Effect of pH on the binding constant (Kb) of kanamycin A (panel A, filled circles) and kanamycin B (panel B, filled circles) to ANT. Determined Δn (Δ) is shown with the overall charge on aminoglycosides ([NH3+]- open squares) and the total charge ([NH3+]+Δn, open diamonds) at three different pH.

The affinity as a function of pH curves for kanamycin A and B binding to ANT show a similar profile from pH 6.4 to 8.1 (Figure 4A and B). At all pH values kanamycin B has a higher affinity for ANT than kanamycin A. Accurate binding data could not be obtained for kanamycin A below pH 6.4 or above 8.1 due to weak binding and the solubility limits of the enzyme. As shown in Figure 4A and B, while the charge ([NH3+]) on the amino groups of kanamycin A, which was determined based on pKas of kanamycin A in solution declines above pH 7.1, an increase in Δn compensates for this loss and the apparent total charge ([NH3+] + Δn) remains unchanged until pH 7.6. Above pH 7.6, both Δn and ([NH3+] + Δn) start to decline. The loss of positive charge on kanamycin parallels the loss of affinity in the pH range 7.6 to 8.1 supporting the conclusion that positively charged amino groups of kanamycin interact with negatively charged amino acids in the active site. Thus loss of such favorable interactions leads to lower affinity. However, the low pH side of the curve clearly shows that this is not the only interaction that affects the affinity of kanamycin A to the enzyme. Deprotonation of a group or groups on the protein, ligand or both decreases the affinity of kanamycin to the protein above pH 6.8. Consistent with the most probable explanation for the binding linked protonation data at pH 7.1, deprotonation of the functional group of an acidic residue (or residues) with an up shifted pKa is likely. Deprotonation that may occur on the aminoglycoside, however, cannot be excluded. Only the 3-amino group of the kanamycin A would be significantly affected in this pH regime (if it remains unaltered or changed only by a small amount upon binding to ANT). If this were the case, the deprotonated form of the 3-NH2 makes more favorable contacts with ANT.

Binding of kanamycin vs. neomycin to ANT

Neomycin is a 4,5-disubstituted aminoglycoside that is a competitive inhibitor of ANT. The pattern of binding linked protonation for this aminoglycoside binding to ANT differs from both kanamycin A and B. In the ANT–neomycin complex, Δn increased continuously with increasing pH (Table 1). At pH 7.1 Δn is slightly negative indicating a small net release of protons from the complex. This value is approximately 0.4 units less than for the ANT–kanamycin B complex. The positions on the A and B ring which are critical for binding ANT are identical in neomycin and kanamycin B [15]. Therefore the difference in binding linked protonation is a result of lowered pKas on either neomycin or ANT caused by interactions between the C or D ring of the antibiotic with the enzyme. A similar Δn is observed for all three antibiotics at pH 7.6. Unlike the kanamycins, however, net proton uptake continues to increase for neomycin from 7.6 to 8.1. This difference may be attributed to the fact that amino groups of neomycin have higher pKa values free in solution and they become even higher when bound to an AGME [24]. Thus, even the highest pH used in this work may be still well below the pKas of its amino groups which causes continued proton uptake.

Another difference between neomycin and the kanamycins is that ΔHint continued to become more negative with increasing pH. In addition, overall change in ΔHint was much larger than those observed for the complexes of kanamycin A and kanamycin B between pH 7.1 and 8.1. The large increase of ΔHint in the ANT–neomycin complex can be accounted for by the ionization of amino groups as reflected in the increasing Δn as pH increases. As indicated earlier, ΔHint may contain contributions from the heat of ionization of titratable groups unless Δn is zero. Intrinsic enthalpy change (ΔHint) determined at a pH where Δn ≈ 0, subtracted from values obtained at regimes where Δn ≠ 0 and the result divided by Δn provides insight into the contribution of ΔHenz to the observed ΔHint. This calculation yields 8.9 and 10.8 kcal mol-1 at pH 7.6 and 8.1 for neomycin. It appears that most of the increase in ΔHint can be attributed to binding-induced pKa shifts of amino groups which typically have a heat of ionization of 9-12 kcal mol-1 [33,34]. A similar calculation with formation of the ANT–kanamycin B complex results in values of 8.3 and 8.5 kcal mol-1 at pH 7.6 and 8.1 respectively. These values are too small for the contribution of amines only and too large for contribution of imidazole (~6 kcal mol-1) indicating that multiple groups with different heats of ionization contribute to ΔHint. Note that the heat of ionization of carboxyl groups (0.5-1 kcal mol-1) would not greatly impact the observed enthalpy change unless several of them contribute to ΔHint. These calculations suggest that either different titratable groups are contributing to ΔHint for kanamycins and neomycin or the same groups contribute differentially due to differences in magnitude of pKa shifts between the two types of antibiotics.

The log Kb versus pH curves for the formation of binary enzyme– aminoglycoside complexes shows that neomycin has a higher affinity for ANT over the entire pH range. At the lower pH range (6.1-7.0) the neomycin and kanamycin B curves are nearly identical (except the neomycin is shifted up at each point). This supports the concept that protonation of the same functional group(s) on the enzyme are responsible for the decreased affinity at pH values less than 6.7. At higher pH values the differences between ANT–neomycin and ANT–kanamycin complexes are more apparent. The affinity for both kanamycin A and B drops sharply above pH 8.2 while only a small decrease in Kb is observed with neomycin (Figures 4 and 5, note the Kb values are presented on a log scale). For neomycin the association constant decreases only from 3.7 x 105 M-1 to 1.4 x 105 M-1. For kanamycin B there is a tenfold decrease in affinity over the same pH range. At pH 8.2, Kb = 1.3 x 105 M-1. At pH 8.4 the association constant decreases to 1.2 x 104 M-1. The difference in the Kb vs. pH profiles at higher pH values can be explained by the continued increase in Δn for neomycin which leads to an increase in. total charge ([NH3 +] + Δn) (Figure 5). This increase is in contrast to kanamycin A and B which decline in total charge above pH 7.6. This is consistent with the presence of groups with up shifted pKas in enzyme–neomycin complex. However the fact that affinity decreases as total charge increases shows that the charge of the amino groups of neomycin cannot solely account for the change in affinity at higher pH. The decrease observed in affinity must be due to loss of other favorable interactions because these groups would still be highly charged.

thermodynamics-catalysis-binding-constant

Figure 5: Effect of pH on the binding constant (Kb) of neomycin (filled symbols) to ANT. Determined Δn (Δ) is shown with the overall charge on aminoglycosides ([NH3+]- open squares) and the total charge ([NH3+]+Δn, open diamonds) at three different pH.

Conclusions

In a system involving multiple ionizable groups on both substrate and enzyme such as aminoglycosides binding to AGMEs, the observed global thermodynamic parameters will reflect contributions from several functional groups. In this work, two important contributors to the pH dependent parameters are identified. First, the importance of the interaction between the 2′ position of the aminoglycoside with the enzyme was determined by observed differences in Δn and Kb between kanamycin A and B. Second, the higher affinity for neomycin compared to kanamycins can be partially attributed to a larger up shift in pKas of amino groups on neomycin. These two factors do not fully account for the pH dependent properties of the ANT–aminoglycoside complexes. The net effect of protonation and deprotonation of other substrate amino groups and enzyme functional groups could not be parsed into individual contributions to the total thermodynamic profile. This difficulty highlights problems with breaking down observed global thermodynamic parameters into individual components, even in simple 1:1 enzyme–substrate complexes. Thus, attribution of changes in these parameters to specific sites should be made carefully since variable and opposing effects may contribute to the same parameter. This is especially true for thermodynamic studies of systems that may include ionizable groups and enzymes that may bind structurally different ligands. Almost all AGMEs are capable of binding structurally different aminoglycosides. In this work properties that contribute to substrate selectivity were shown by determining binding linked protonation and affinity as a function of pH. Structural and further thermodynamic analysis will be necessary for a more detailed analysis of the factors that contribute to the substrate selectivity and promiscuity of ANT.

Acknowledgement

This research was partly supported by a Grant from the National Science Foundation (MCB 0842743 to EHS).

References

  1. Ramirez MS, Tolmasky ME (2010) Aminoglycoside modifying enzymes. Drug Resist Updat 13: 151-171.
  2. Shaw KJ, Rather PN, Hare RS, Miller GH (1993) Molecular genetics of aminoglycoside resistance genes and familial relationships of the aminoglycoside-modifying enzymes. Microbiol Rev 57: 138-163.
  3. Benveniste R, Davies J (1973) Mechanisms of antibiotic resistance in bacteria. Annu Rev Biochem 42: 471-506.
  4. O'Brien TF, Pla MP, Mayer KH, Kishi H, Gilleece E, et al. (1985) Intercontinental spread of a new antibiotic resistance gene on an epidemic plasmid. Science 230: 87-88.
  5. Shimizu K, Kumada T, Hsieh WC, Chung HY, Chong Y, et al. (1985) Comparison of aminoglycoside resistance patterns in Japan, Formosa, and Korea, Chile, and the United States. Antimicrob Agents Chemother 28: 282-288.
  6. Poole K (2005) Aminoglycoside resistance in Pseudomonas aeruginosa. Antimicrob Agents Chemother 49: 479-487.
  7. Strateva T, Yordanov D (2009) Pseudomonas aeruginosa - a phenomenon of bacterial resistance. J Med Microbiol 58: 1133-1148.
  8. Bongaerts GP, Molendijk L (1984) Relation between aminoglycoside 2"-O-nucleotidyltransferase activity and aminoglycoside resistance. Antimicrob Agents Chemother 25: 234-237.
  9. Smith AL, Smith DH (1974) Gentamicin:adenine mononucleotide transferase: partial purification, characterization, and use in the clinical quantitation of gentamicin. J Infect Dis 129: 391-401.
  10. Wright E, Serpersu EH (2004) Isolation of aminoglycoside nucleotidyltransferase(2")-Ia from inclusion bodies as active, monomeric enzyme. Protein Expr Purif 35: 373-380.
  11. De Hertogh DA, Lerner SA (1985) Correlation of Aminoglycoside Resistance with the Kms and Vmax/Km Ratios of Enzymatic Modification of Aminoglycosides by 2''-O-Nucleotidyltransferase. Antimicrob Agents Chemother 27: 670-671.
  12. Gates CA, Northrop DB (1988) Substrate specificities and structure-activity relationships for the nucleotidylation of antibiotics catalyzed by aminoglycoside nucleotidyltransferase 2''-I. Biochemistry 27: 3820-3825.
  13. Van Pelt JE, Northrop DB (1984) Purification and Properties of Gentamicin Nucleotidyltransferase from Escherichia-Coli - Nucleotide Specificity, Ph Optimum, and the Separation of 2 Electrophoretic Variants. Arch Biochem Biophys 230: 250-263.
  14. Wright E, Serpersu EH (2005) Enzyme-substrate interactions with an antibiotic resistance enzyme: Aminoglycoside nucleotidyltransferase(2 '')-Ia characterized by kinetic and thermodynamic methods. Biochemistry 44: 11581- 11591.
  15. Wright E, Serpersu EH (2006) Molecular determinants of affinity for aminoglycoside binding to the aminoglycoside nucleotidyltransferase(2 '')-Ia. Biochemistry 45: 10243-10250.
  16. Burkhalter B, Dimick S, Toone E (2000) Protein-Carbohydrate Interaction: Fundamental Considerations, in Carbohydrates in Chemistry and Biology (Ernst, B., GWH, and Sinay, P., Eds.) pp 863-914, Wiley-VCH, New York.
  17. Chervenak MC, Toone EJ (1995) Calorimetric Analysis of the Binding of Lectins with Overlapping Carbohydrate-Binding Ligand Specificities. Biochemistry 34: 5685-5695.
  18. Dam TK, Brewer CF (2002) Thermodynamic studies of lectin-carbohydrate interactions by isothermal titration calorimetry. Chem Rev 102: 387-429.
  19. Norris AL, Ozen C, Serpersu EH (2010) Thermodynamics and Kinetics of Association of Antibiotics with the Aminoglycoside Acetyltransferase (3)-IIIb, a Resistance-Causing Enzyme. Biochemistry 49: 4027-4035.
  20. Ozen C, Serpersu EH (2004) Thermodynamics of aminoglycoside binding to aminoglycoside-3 '-phosphotransferase IIIa studied by isothermal titration calorimetry. Biochemistry 43: 14667-14675.
  21. Goldberg RN, Kishore N, Lennen RM (2002) Thermodynamic Quantities for the Ionization Reactions of Buffers. J Phys Chem Ref Data 31: 231-370.
  22. Wiseman T, Williston S, Brandts JF, Lin LN (1989) Rapid Measurement of Binding Constants and Heats of Binding Using a New Titration Calorimeter. Anal Biochem 179: 131-137.
  23. Atha DH, Ackers GK (1974) Calorimetric Determination of Heat of Oxygenation of Human Hemoglobin as a Function of pH and Extent of Reaction. Biochemistry 13: 2376-2383.
  24. Ozen C, Malek JM, Serpersu EH (2006) Dissection of aminoglycoside-enzyme interactions: A calorimetric and NMR study of neomycin B binding to the aminoglycoside phosphotransferase(3')-IIIa. J Am Chem Soc 128: 15248- 15254.
  25. Young PG, Walanj R, Lakshmi V, Byrnes LJ, Metcalf P, et al. (2009) The Crystal Structures of Substrate and Nucleotide Complexes of Enterococcus faecium Aminoglycoside-2 ''-Phosphotransferase-IIa [APH(2 '')-IIa] Provide Insights into Substrate Selectivity in the APH(2 '') Subfamily. J Bacteriol 191: 4133-4143.
  26. Fong DH, Berghuis AM (2002) Substrate promiscuity of an aminoglycoside antibiotic resistance enzyme via target mimicry. EMBO J 21: 2323-2331.
  27. Vetting MW, Magnet S, Nieves E. Roderick SL, Blanchard JS (2004) A bacterial acetyltransferase capable of regioselective N-acetylation of antibiotics and histones. Chem Biol 11: 565-573.
  28. Pedersen LC, Benning MM, Holden HM (1995) Structural Investigation of the Antibiotic and ATP-Binding Sites in Kanamycin Nucleotidyltransferase. Biochemistry 34: 13305-13311.
  29. Harris TK, Turner GJ (2002) Structural basis of perturbed pKa values of catalytic groups in enzyme active sites. IUBMB Life 53: 85-98.
  30. Honig B, Nicholls A (1995) Classical Electrostatics in Biology and Chemistry. Science 268: 1144-1149.
  31. Li H, Robertson AD, Jensen JH (2004) The determinants of carboxyl pKa values in Turkey ovomucoid third domain. Proteins 55: 689-704.
  32. Boehr DD, Draker KA, Koteva K, Bains M, Hancock RE, et al. (2003) Broadspectrum peptide inhibitors of aminoglycoside antibotic resistance enzymes. Chem Biol 10: 189-196.
  33. Marini MA, Berger RL, Lam DP, Martin CJ (1971) Thermal titrimetric evaluation of the heats of ionization of the commonly occurring amino acids and their derivatives. Anal Biochem 43: 188-198.
  34. Bjurulf C (1972) Calorimetric titration of lysozyme. Eur J Biochem 30: 33-36.
Citation: Wright E, Serpersu EH (2011) Effects of Proton Linkage on Thermodynamic Properties of Enzyme-Antibiotic Complexes of the Aminoglycoside Nucleotidyltransferase (2")-Ia. J Thermodyn Catal 2: 105.

Copyright: © 2011 Wright E, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top