Next Article in Journal
Glycerosome of Melissa officinalis L. Essential Oil for Effective Anti-HSV Type 1
Next Article in Special Issue
Synthesis of Water-Soluble Copolymers of N-vinylpyrrolidone with N-vinyldithiocarbamate as Multidentate Polymeric Chelation Systems and Their Complexes with Indium and Gallium
Previous Article in Journal
Synthetic Mono-Rhamnolipids Display Direct Antifungal Effects and Trigger an Innate Immune Response in Tomato against Botrytis Cinerea
Previous Article in Special Issue
A New Tetradentate Mixed Aza-Thioether Macrocycle and Its Complexation Behavior towards Fe(II), Ni(II) and Cu(II) Ions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interaction of Ampicillin and Amoxicillin with Mn2+: A Speciation Study in Aqueous Solution

Dipartimento di Scienze Chimiche, Biologiche, Farmaceutiche ed Ambientali, Università di Messina, Viale F. Stagno d’Alcontres 31, 98166 Messina, Italy
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(14), 3110; https://doi.org/10.3390/molecules25143110
Submission received: 15 June 2020 / Revised: 3 July 2020 / Accepted: 6 July 2020 / Published: 8 July 2020
(This article belongs to the Special Issue Chelating Agents towards Bioavailable Metal Ions)

Abstract

:
A potentiometric and UV spectrophotometric investigation on Mn2+-ampicillin and Mn2+-amoxicillin systems in NaCl aqueous solution is reported. The potentiometric measurements were carried out under different conditions of temperature (15 ≤ t/°C ≤ 37). The obtained speciation pattern includes two species for both the investigated systems. More in detail, for system containing ampicillin MLH and ML species, for that containing amoxicillin, MLH2 and MLH ones. The spectrophotometric findings have fully confirmed the results obtained by potentiometry for both the systems, in terms of speciation models as well as the stability constants of the formed species. Enthalpy change values were calculated via the dependence of formation constants of the species on temperature. The sequestering ability of ampicillin and amoxicillin towards Mn2+ was also evaluated under different conditions of pH and temperature via pL0.5 empirical parameter (i.e., cologarithm of the ligand concentration required to sequester 50% of the metal ion present in traces).

Graphical Abstract

1. Introduction

Photosynthesis, proteoglycan biosynthesis and antioxidative actions are the main biochemical functions [1]. Mn2+ is the most stable thermodynamically oxidation state in both acidic and basic solutions [2,3,4]. In biological systems Mn2+ can be substituted and can substitute other divalent cations, such as Ca2+, Mg2+, Fe2+. Mn2+ and Mg2+ may exchange cofactor roles in enzyme reactions and they show the same affinity for simple oxygen donors, such as polyphosphates [5,6]. The main difference among Mn(II) and Mg(II) chemistry consists in the higher affinity of manganese ion for nitrogen or sulphur ligands. It presents an intermediate size and an intermediate electron acceptor strength between Ca(II) and Mg(II). In biological systems, Mn(II) concentration is about 0.1 μmol L−1 inside and outside cells [6]. High manganese concentrations have been detected in the pancreas, liver, kidney and intestines, where the manganese content was found in the range 2.5–5.3 and 4.9–6.0 μg kg−1, in kidney and liver respectively [1]. In blood, manganese is bound to plasma proteins, its concentration in the whole blood is about 11 μg L−1 and in the serum 0.6–1.3 μg L−1. In the urine the mean concentration of 9.32 μg L−1 was found [7]. For humans the main sources of manganese are foods and beverages and its absorption occurs mainly through respiratory and gastrointestinal tracts [1]. In humans, animals and plants, manganese is an activator as well as a constituent of various enzymes [5]. As a consequence of its excess or deficiency numerous pathologies can occur [8]. Thereby, it acts as both an essential nutrient and toxic element. Under normal conditions, the body protects itself through the homeostatic mechanism, acting on absorption and excretion [1]. In humans, symptoms of both manganese deficiency and oral parenteral poisoning are rare [9]. Manganese is also widely present in waters, soils, sediments, air. Its concentration in the open oceans, mainly as Mn2+, range between 0.01 and 0.16 μg kg−1. More in detail, manganese concentration is in the range 1–500 μg L−1 in surface waters, 5–25 μg L−1 in drinking water [1]. The simple [Mn(H2O)6]2+ ion is commonly found in freshwaters and in seawaters [4]. Similarly to the other bivalent cations of the transition series, Mn2+ can be considered resistant to hydrolysis [2]. At ordinary concentrations, hydrolysis starts to be significant above pH 8 [2]. The interactions of Mn2+ with antibiotics are of fundamental importance both from biological and environmental point of view. The biological interest is more obvious than the environmental one, however, one must consider that antibiotics are excreted via feces and urine, through domestic and hospital wastewaters. They are thus released into aquatic environments, soils, sediments and many of them persist for long time in the environment [10]. In this context, it is crucial the study of their interactions with metal cations, in order to evaluate the mechanisms by which they can be retained in natural waters.
Among antibiotics, the aminopenicillins are semi-synthetic penicillins that can be considered as N-acylderivatives of 6-aminopenicillanic acid. Semi-synthetic β-lactam antibiotics have been designed for a wider spectrum of action and better pharmacokinetic properties than natural penicillin. Among semi-synthetic penicillins, ampicillin and amoxicillin have the best oral absorption and they distribute well in body fluids. They are commonly employed as antibiotics owing to the wide spectra mechanisms of action, since they stop bacteria proliferation [11]. It is known that the bioavailability of β-lactam antibiotics is influenced by their interaction in vivo with several metal cations [12,13,14,15,16,17]. For this reason, the assessment of the interactions of these antibiotics with metal cations which play biological roles in the human body is crucial [18]. Complexing abilities of the most important amino-penicillins, i.e., ampicillin and amoxicillin, with some metal cations, namely Zn2+, Cu2+, Ca2+ and Mg2+, were already studied [19,20,21,22].
In this paper the interactions of Mn2+ with ampicillin (Amp) and amoxicillin (Amox), shown in Figure 1, were investigated. With the aim to clarify the in vivo mode of action of these antibiotics with a bioavailable metal cation, the study of the interactions of Amp and Amox with Mn2+ was performed [23]. The investigation on these systems was carried out by potentiometry and UV spectrophotometry. Potentiometric titrations were performed under different conditions of ionic strengths and temperature in order to model the dependence of formation constants on temperature and on ionic strength. Spectrophotometric titrations were used to confirm the speciation models and the stability of complex species.
Sequestering ability of Amp and Amox was quantified via a Boltzmann type equation where the sum of molar fractions of all Mn2+-ligand complex species were fitted vs. cologarithm of total ligand concentration. Sequestering ability under different conditions of temperature, pH and ionic strength was also evaluated.

2. Results and Discussion

2.1. Aqueous Behaviour and Speciation of Mn2+-Amp and Mn2+-Amox Species

In this investigation, potentiometric and UV spectrophotometric titrations were performed in order to define the best speciation models for both the investigated systems and to obtain reliable formation constants values of the complex species. Formation equilibria of metal cation(M)-ligand(L) species are the following (were the charges are omitted for simplicity):
M + L + rH = MLHr      βr
M + LHr = MLHr       Kr
Hydrolysis constants of Mn2+ and protonation constants of Amp and Amox, were taken into account in the calculations (Tables S1 and S2, respectively of Supplemetary Materials). The elaboration of the potentiometric experimental data has allowed to obtain the results listed in Table 1. The choice of the best speciation model is always a very delicate phase in the study of the equilibria in solution. It was chosen on the basis of a series of fundamental criteria, such as the best statistical fit, the variance ratio between the accepted model and the others, the model simplicity, the formation percentages of the complex species [24,25]. Taking into account what has just been described, the speciation model chosen for the Mn2+(M)-Amp(L) system, under all the temperature and ionic strength conditions investigated, was found to contain only 1:1 M:L species, namely MLH and ML.
Figure 2 shows the speciation diagram for this system at t = 25 °C, CM = 2 mmol L−1, CL = 4 mmol L−1, I = 0.15 mol L−1 in NaCl. It can be observed from the diagram that the MLH species is present in a wide pH range with metal fraction greater than 0.4, the ML one reaches almost 0.4 as a fraction at pH = 8.5. The M2(OH)3 hydrolytic species is significant only starting from pH > 9.
In order to confirm the speciation model and the formation constants gained via potentiometry by spectroscopic techniques as already established for other systems [26,27,28,29,30,31,32,33,34], spectrophotometric titrations on solutions containing Mn2+ and Amp at t = 25 °C and I = 0.15 mol L−1 were performed under different concentration conditions. As an example, experimental spectra recorded on solutions containing Mn2+ and Amp at CM = 0.05 mmol L−1 and CL = 0.05 mmol L−1, were depicted in Figure 3. It shows that the absorption spectra have a single maximum with a slight hypochromic and hypsochromic effect with increasing pH.
Via the processing of the spectrophotometric experimental data with the HypSpec program the results listed in Table 2 were obtained, together with the potentiometric ones under the same conditions of temperature and ionic strength. An examination of both results shows that the speciation model proposed by potentiometry was confirmed by spectrophotometry and the formation constants of both complex species were obtained with very close values to those gained by potentiometry. The molar absorbances of Amp species and of Mn2+-Amp complexes at t = 25 °C and I = 0.15 mol L−1 in NaCl are represented in Figure 4. It is observed that both MLH and ML complex species have molar absorbances higher than those of the single L, LH, LH2 species of Amp.
The processing of the potentiometric experimental data carried out on Mn2+-Amox solutions has provided the results listed in Table 1. The speciation model obtained for Mn2+(M)-Amox(L) system, under all the investigated ionic strength and temperature conditions, includes only complex species with 1:1 M:L ratio, i.e., MLH2 and MLH. The speciation diagram regarding this system at CM = 2 mmol L−1, CL = 4 mmol L−1, t = 25 °C, I = 0.15 mol L−1 in NaCl is depicted in Figure 5. It is observed that MLH2 species predominates in the acid pH range, 3 ≤ pH ≤ 6.5, with metal fractions close to 0.6. MLH is the main species in the range 7 ≤ pH ≤ 9, with a maximum metal fraction almost equal to 0.8 at pH = 8.5. M2(OH)3 hydrolytic species does not reach significant formation percentages up to pH = 9.
As for Amp, the system containing Mn2+-Amox has been investigated with potentiometry as well as with UV spectrophotometry. Experimental UV spectra at different pH values recorded on solutions containing Mn2+(M) and Amox(L) are shown in Figure 6 (CM = 0.05 mmol L−1, CL = 0.075 mmol L−1, t = 25 °C, I = 0.15 mol L−1 in NaCl). As known, spectra of solutions containing Amox result more complex than those containing Amp. In the range λ = 220–310 nm the former present intense bands, as expected for ligands having heteroatoms, linked to benzene ring, giving electronic transfer phenomena. Accordingly, the observed absorptions are typical of aromatic compounds with polar substituents, such as phenolic groups, characterized by charge transfer with ε = 5 × 103–16 × 103 L mol−1 cm−1 in the range mentioned above.
The UV spectra in terms of absorptivity, shape and band position are influenced by alkyl substitution, which generally causes a bathochromic shift of the phenolic band [35]. The presence of Mn2+ causes a slight hypsocromic shift of the phenolic band which can be explained by admitting metal-ligand interaction. More in detail, in the UV spectra shown in Figure 6 there are three maxima in the range of pH = 5–9.5, at λ = 204, 229 and 272 nm, at pH = 10.5 the maxima are at λ = 207, 248, 289 nm.
These absorption maxima show different trends with increasing pH, the maxima at about 204 and 272 nm undergo a hyperchromic effect, on the contrary the maximum at 229 nm has a hypochromic effect. Molar absorbances of Mn2+-Amox species as well as those of Amox are depicted in Figure 7. More in detail, at λ = 207 nm, fully deprotonated ligand species has ε ≅ 27 × 103 L mol−1 cm−1, followed by MLH and MLH2 with ε ≅ 21 × 103, 18 × 103 L mol−1 cm−1, respectively at λ = 204 nm. At λ = 229 nm, only ligand protonated species and complex ones present a maximum. LH2 species shows ε ≅ 1 × 103 L mol−1 cm−1 at 229 nm, the interaction with metal causes a slight shift of the maximum. Accordingly, MLH2 and MLH present a maximum at 228 nm with ε ≅ 9 × 103.
The results obtained by processing the spectrophotometric experimental data are listed in Table 2 together with the potentiometric ones under the same conditions of ionic strength and temperature. From the comparison, it is possible to affirm that spectrophotometry confirmed the same speciation model gained by potentiometry and the formation constants of both complex species were calculated with values quite similar to potentiometric ones.

2.2. Dependence of Formation Constants on Ionic Strength

The experimental results of Table 1, at various ionic strength values, were analyzed considering the Debye-Hückel equation:
log β = log β 0 0.51 z * I 1 + 1.5 I + C I
where β0 is the formation constant at infinite dilution, z* = ∑(charge)2reactants − ∑(charge)2products, C is an empirical parameter. Several examples of these applications are reported in previous papers [36,37,38,39,40]. Calculated values of the formation constants at infinite dilution as well as of the C parameter for the dependence on the ionic strength are listed in Table 3. These parameters, necessary to calculate the stability constants at other ionic strengths, are of great importance for applications to real systems, such as biological fluids or natural waters. Recalculated values are listed in Table S3 of Supplemtary Materials.
By using the experimental results of Table 1, the speciation diagrams at I = 0.15 and 1 mol L−1 were obtained, as depicted in Figure 2 and Figure 5.

2.3. Dependence of Formation Constants on Temperature

Potentiometric titrations were performed at t = 15, 25, 37 °C and I = 0.15 mol L−1 for both Mn2+-Amp and Mn2+-Amox systems. The formation constant values at the various temperatures investigated are listed in Table 1, for both the systems. These data were analyzed using the van’t Hoff equation:
logβT = logβθ + ΔH0(1/θ − 1/T) Rln10
where logβT is the formation constant at a temperature T expressed in Kelvin, logβθ is the formation constant at T = 298.15 K, ΔH0 is expressed in kJ mol−1, R = 8.314472 J K−1 mol−1.
Enthalpy change values relating to all species of Mn2+-Amp and Mn2+-Amox systems were determined by using Equation (4). These values, together with those of free energy and entropy changes, are reported in Table 4, referring both on global (1) and on partial reaction (2). For Mn2+-Amp complexes ΔH values are exothermic for both the species. The trend of the thermodynamic parameters of Mn2+-Amp species is the same of Mg2+, the cation having the most similar biological behaviour to Mn2+, More in detail, for ML species of Mg2+-Amp system ΔH = −31.1 kJ mol−1, TΔS = −12.6 kJ mol−1 (at t = 25 °C and I = 0 mol L−1), fairly similar to ΔH = −21 kJ mol−1, TΔS = −8 kJ mol−1 here reported for the same species containing Mn2+.
For Mn2+-Amox complexes, by considering partial reaction (2), ΔH value is exothermic for MLH species and endothermic for MLH2. Referring to partial reactions, the prevalence of the enthalpy and the entropic contribution for the species containing Amp and Amox, respectively, is not marked at all. Figure 8, represents a bar plot for both the systems which evidences the trend of thermodynamic parameters.
Figure 9 and Figure 10 show the speciation diagrams at t = 15, 37 °C relating to Mn2+-Amp and Mn2+-Amox systems, respectively. In both systems the change in temperature leads very different formation percentages of complex species, highlighting the need to take into account this parameter for a correct speciation study.

2.4. Sequestering Ability

Often the simple comparison of the formation constant values and the formation percentages of the species is not sufficient to analyze the higher or lesser sequestering capacity of a ligand towards different metal cations or of different ligands towards the same metal cation. In order to assess the sequestering ability, it is necessary to consider all the equilibria in which the ligands and various metal cations participate, such as hydrolysis reactions, ligand protonations, weak interactions of the ligand with the cations of the background salt. Several years ago, an empirical parameter, pL0.5, widely tested on multiple metal-ligand systems, was proposed for this purpose. pL0.5 represents the cologarithm of the concentration of the ligand able to bind 50% of the metal cation in traces, under the investigated conditions of pH, temperature and ionic strength. The following Boltzmann equation with asymptotes of 0 for pL→0 and 1 for pL→∞ is used to quantitatively define the sequestering capacity of a ligand against a specific metal cation [41,42,43,44,45]:
χ = 1 1 + 10 ( p L p L 0.5 )
where χ is the sum of the molar fractions of the species and pL represents the cologarithm of the total ligand concentration. In this way, the whole composition of the system is taken into account by pL0.5, as it depends on the system conditions (e.g., pH, temperature, ionic strength), considering all competitive reactions of the metal cation and the ligand. Moreover, in the calculation of the sequestering capacity a concentration of metal cation in traces was considered, making this parameter closer to the concentration conditions of many metal cations in real systems, such as body fluids. Under physiological conditions (pH = 7.4, t = 37 °C, I = 0.15 mol L−1), the sequestering ability values calculated against Mn2+ for Amp and Amox resulted in pL0.5 = 2.01, 3.17, respectively. It emerges that under physiological conditions the binding capacity of Amox is more than one logarithmic unit compared to Amp. These results are comparable with those obtained with another divalent cation, Ca2+, under the same conditions, pL0.5 = 1.82, 2.88 for Amp and Amox, respectively [22]. Another comparison can be made with Zn2+, in this case in the same conditions the sequestering ability is comparable for Amox, while for Amp it is significantly different. More in detail, pL0.5 = 3.16, 2.88 for Amp and Amox, respectively, towards Zn2+ [19]. The sequestering ability values of Amp and Amox towards Mn2+ were calculated under different conditions of pH, temperature and ionic strength. These values are listed in Table 5. Figure 11 shows the comparison of the sequestering capacity of the two ligands towards Mn2+ under physiological conditions. Binding ability of Amox is significantly higher than Amp under these conditions.

2.5. Literature Comparisons

In literature, in the main databases that report thermodynamic data [46,47,48], there are no formation constant values for Mn2+-Amp and Mn2+-Amox species. Only formation constant values for Cu2+, Ni2+, Zn2+, Cd2+, Co2+ are found. As far as we know, no thermodynamic data are reported on the complex species formed by Amox with any metal cation, apart from the results already published on Ca2+, Mg2+, Cu2+, Zn2+ cations [19,20,21,22]. Therefore, it is possible to compare simply the results here reported on Mn2+ with those obtained with other divalent cations. Among these cations, the behaviour of Mn2+ is quite similar to Mg2+ and Ca2+. More in detail, for the system containing Amp, formation constant values for MgLH and MgL species are 2.54, 2.86, respectively (at I = 0.15 mol L−1 and t = 25 °C) [21], for the same species containing Mn2+ are 2.47, 2.37, respectively, as reported here under the same conditions. For the system containing Amox, formation constant values for MLH and ML species with Ca2+ are 2.69, 2.59, respectively (at I = 0.15 mol L−1 and t = 25 °C) [22], for the same species with Mn2+ are 2.78, 3.20, respectively, as reported here under the same conditions.

3. Materials and Methods

3.1. Materials

The solutions containing Mn2+ were prepared by weighing and dissolution of the corresponding salt, manganese(II) chloride tetrahydrate (Reagent Plus ≥ 99%, Sigma, Darmstadt, Germany) These solutions were standardized by titration with ethylenediamine tetraacetic acid disodium salt (EDTA) standard solution. Solutions containing ligand were obtained by weighing and dissolution of ampicillin anhydrous and amoxicillin trihydrate (analytical standard, Sigma-Aldrich), used without further purification. Ligand purity, checked by alkalimetric titration, resulted > 99%. Fresh solutions of ligands were prepared every day. Hydrochloric acid and sodium hydroxide solutions were prepared by dilution of the respective ampoules (Fluka, Munich, Germany) and after they were standardized with sodium carbonate (≥99.5%, Sigma-Aldrich) and potassium biphthalate (≥99.5%, Sigma-Aldrich), respectively, previously dried in an oven at 110 °C. Sodium hydroxide solutions were prepared very frequently and were stored in bottles containing soda lime traps. Sodium chloride solutions were obtained by weighing the salt (Sigma-Aldrich®, puriss.), previously dried at 110 °C. All the solutions were prepared using bidistilled water (conductivity < 0.1 μS cm−1) and grade A glassware.

3.2. Potentiometric Apparatus and Procedure

Potentiometric titrations were performed by two distinct systems including a Metrohm (Herisau, Switzerland)) model 809 Titrando potentiometer, a Metrohm LL-Unitrode WOC combined glass electrode, and an automatic dispenser Metrohm Dosino 800. The system just described was connected to a PC and the measurements were acquired by the Metrohm TIAMO 2.2 software, which can control titrant delivery, e.m.f. stability, and data acquisition. Estimated accuracy of this system is ±0.15 mV and ±0.002 mL for e.m.f. and for titrant volume readings, respectively.
During the titration NaOH standard was added to 25 mL of the solution containing Mn2+, ligand and a supporting electrolyte (NaCl). For Mn2+(M)-Amp(L) solutions, 1 ≤ CM/mmol L−1 ≤ 3, 1 ≤ CL/mmol L−1 ≤ 6 and concentration ratios 1 ≤ CL/CM ≤ 3. For Mn2+(M)-Amox(L) solutions, 1.5 ≤ CM/mmol L−1 ≤ 2, 2 ≤ CL/mmol L−1 ≤ 4 and concentration ratios 1 ≤ CL/CM ≤ 2. The investigated pH ranges between 2 and 10. The measurements were performed into glass jacket thermostated cells, under different conditions of temperature (15 ≤ t/°C ≤ 37) and ionic strength (0.15 ≤ I/mol L−1 ≤ 1.0), under magnetic stirring and by bubbling pure N2 in order to avoid CO2 and O2 inside the solutions. An independent titration of HCl with standard NaOH was carried out for each measurement to calculate the pKw value and the standard electrode potential E0 under the same experimental conditions of ionic strength and temperature of the measurement itself.

3.3. UV-Vis Apparatus and Procedure

A Varian (Agilent Scientific Instruments (Santa Clara, CA, USA)) Cary 50 UV-Vis spectrophotometer equipped with an optical fiber probe having a path length equal to 1 cm, was employed for spectrophotometric titrations on Mn2+-Amp and Mn2+-Amox solutions. The instrument was connected to a PC for the acquisition of the UV spectra in the range from 200 to 320 nm. A volume of 25 mL of the solution containing the metal cation and the ligand at different metal/ligand ratios was investigated at t = 25 °C, I = 0.15 mol L−1 in NaCl and at different pH values. For Mn2+(M)-Amp(L) solutions, 0.04 ≤ CM/mmol L−1 ≤ 0.05, 0.05 ≤ CL/mmol L−1 ≤ 0.08 and concentration ratios 1 ≤ CL/CM ≤ 2. For Mn2+(M)-Amox(L) solutions, 0.05 ≤ CM/mmol L−1 ≤ 0.1, 0.075 ≤ CL/mmol L−1 ≤ 0.12 and concentration ratios 1 ≤ CL/CM ≤ 2. The titrations were performed by potentiometric apparatus already described in the 3.2 paragraph. Aliquots of standard NaOH were added to solution containing Mn2+-Amp and Mn2+-Amox in order to reach the maximum absorbance of each complex species.

3.4. Calculations

Experimental potentiometric data were processed via the BSTAC and STACO programs, in order to gain the best speciation model for the systems under study, the formation constant values of the complex species, and all the parameters of an acid-base titration (analytical concentration of the reagents, standard potential E0, junction potential). LIANA software was employed in order to obtain the parameters for the dependence of complex formation constants on temperature and ionic strength. Details on software used in the refinement of the experimental data are reported in ref. [49]. Experimental spectrophotometric data were refined by HypSpec program to calculate the molar absorbance spectrum and the stability constant values of complex species [50,51]. HySS program was employed to obtain the speciation diagrams and the formation percentages of the complex species [52].

4. Conclusions

Due to the widespread use of aminopenicillins as antibiotics, knowledge of their complexation behavior and speciation is of fundamental importance for understanding their pharmacokinetics and pharmacodynamics. Speciation studies on Mn2+-Amp and Amox solutions were performed by potentiometry under different experimental conditions of temperature and ionic strengths. The results in terms of most reliable speciation model and formation constant values were fully confirmed by UV spectrophotometry for both the systems. More in detail, despite the different metal and ligand concentrations employed with these two techniques, a satisfactory agreement among the stability constants was achieved. The dependence of the stability constants on temperature and on ionic strength was studied. Accordingly, enthalpy changes at t = 25 °C and parameters for the dependence on ionic strength were also calculated.
For the correct assessment of the bioavailability of a drug or a trace element in the presence of that drug, it is essential to know the possible interactions occurring in body fluids. Accordingly, the knowledge of speciation and sequestering ability of two drugs employed as antibiotics, such as Amp and Amox with Mn2+, are necessary to evaluate the action mechanism of these ligands and the bioavailability of the metal cation in the presence of the drugs under physiological conditions. The assessment of the sequestering capacity is of fundamental importance as it takes into account the competitive processes of the metal cation and the ligand and facilitates comparisons among several ligands towards the same metal cation or the same ligand towards several cations under the same conditions of pH, temperature and ionic strength. The sequestering ability of Amp and Amox towards Mn2+ was evaluated under different conditions, including physiological ones. Under these conditions Amox has shown a higher binding capacity towards Mn2+ of more than one logarithmic unit compared to Amp.

Supplementary Materials

The following are available online, Table S1: Hydrolysis constant values of Mn2+ at different temperatures and ionic strengths in NaCl, Table S2: Protonation constant values of Amp and Amox at different temperatures and ionic strengths in NaCl, Table S3: Recalculated formation constant values for Mn2+-Amp and Mn2+-Amox species at different ionic strengths and temperatures.

Author Contributions

Conceptualization, O.G. and C.F.; methodology, O.G.; validation, O.G. and C.F.; investigation, O.G; data curation, O.G.; writing—original draft preparation, O.G.; writing—review and editing, O.G. and C.F.; supervision, O.G.; project administration, O.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schafer, U. Manganese. In Elements and their Compounds in the Environment, 2nd ed.; Merian, E., Anke, M., Ihnat, M., Stoeppler, M., Eds.; Wiley-VCH Verlag GmbH & Co.: Weinheim, Germany, 2004; pp. 901–930. [Google Scholar]
  2. Baes, C.F.; Mesmer, R.E. The Hydrolysis of Cations; John Wiley & Sons: New York, NY, USA, 1976. [Google Scholar]
  3. Reidies, A.H. Manganese Compounds. In Ulmann’s Encyclopedia of Industrial Chemistry, 6th ed.; Wiley-VCH Verlag: Weinheim, Germany, 2003; Volume 20, pp. 543–564. [Google Scholar]
  4. Chiswell, B.; Johnson, D. Manganese. In Handbook on Metals in Clinical and Analytical Chemistry; Seiler, H.G., Sigel, A., Sigel, H., Eds.; Marcel Dekker, Inc.: New York, NY, USA; Basel, Switzerland; Hong Kong, China, 1994. [Google Scholar]
  5. Morgan, J.J. Manganese in Natural Waters and Earth’s Crust: Its Availability to Organisms. In Manganese and Its Role in Biological Systems; Sigel, A., Sigel, H., Eds.; Marcel Dekker, Inc.: New York, NY, USA; Basel, Switzerland, 2000; Volume 37, pp. 1–34. [Google Scholar]
  6. Frausto da Silva, J.J.R.; Williams, R.J.P. Manganese: Dioxygen evolution and glycosylation. In The Biological Chemistry of the Elements: The Inorganic Chemistry of Life, 2nd ed.; Oxford University Press: Oxford, UK, 2001. [Google Scholar]
  7. Schafer, U.; Anke, M.; Seifert, M.; Fischer, A.B. Influences on the manganese intake, excretion and balance of adults, and on the manganese concentration of the consumed food determined by means of the duplicate portion technique. Trace Elem. Electrolytes 2004, 21, 68–77. [Google Scholar] [CrossRef]
  8. Kabata-Pendias, A.; Pendias, H. Trace Elements in Soils and Plants, 3rd ed.; CRC Press: Boca Raton, FL, USA; London, UK; New York, NY, USA; Washington, DC, USA, 2001. [Google Scholar]
  9. Kondakis, X.G.; Makris, N.; Leotsinidis, M.; Prinou, M.; Papapetropoulos, T. Possible Health Effects of High Manganese Concentration in Drinking Water. Arch. Env. Health 1989, 44, 175–178. [Google Scholar] [CrossRef]
  10. Bouyarmane, H.; El Hanbali, l.; El Karbane, M.; Rami, A.; Saoiabi, A.; Saoiabi, S.; Masse, S.; Coradin, T.; Laghzizil, A. Parameters influencing ciprofloxacin, ofloxacin, amoxicillin and sulfamethoxazole retention by natural and converted calcium phosphates. J. Hazard. Mater. 2015, 291, 38–44. [Google Scholar] [CrossRef] [Green Version]
  11. Proctor, P.; Gensmantel, N.P.; Page, M.I. The chemical reactivity of penicillins and other β-lactam antibiotics. J. Chem. Soc. Perkin Trans. 1982, 1185–1192. [Google Scholar] [CrossRef]
  12. Imran, M.; Iqbal, J.; Mehmood, T.; Latif, F. Synthesis, characterization and in vitro screening of amoxicillin and its complexes with Ag(I), Cu(II), Co(II), Zn(II) and Ni(II). J. Biol. Sci. 2006, 6, 946–949. [Google Scholar]
  13. Gutiérez Navarro, P.; El Bekkouri, A.; Rodriguez Reinoso, E. Spectrofluorimetric study of the degradation of [small alpha]-amino [small beta]-lactam antibiotics catalysed by metal ions in methanol. Analyst 1998, 123, 2263–2266. [Google Scholar] [CrossRef] [PubMed]
  14. Gutiérrez Navarro, P.; Hernandez Blasquez, I.; Quintero Osso, B.; Martinez de las Parras, P.J.; Martinez Puentedura, M.I.; Marquez Garcia, A.A. Penicillin degradation catalysed by Zn(II) ions in methanol. Int. J. Biol. Macromol. 2003, 33, 159–166. [Google Scholar] [CrossRef]
  15. Marquez Garcia, A.; Gutiérrez Navarro, P.; Martinez de las Parras, P.J. Degradation of ampicillin in the presence of cadmium(II) ions. Talanta 1998, 46, 101–109. [Google Scholar] [CrossRef] [Green Version]
  16. Sher, A.; Veber, M.; Marolt-Gomiscek, M.; Gomiscek, S. Complexation of copper(II) ions with ampicillin I: Spectroscopic and electrochemical investigation of interactions under equilibrium conditions. Int. J. Pharm. 1993, 90, 181–186. [Google Scholar] [CrossRef]
  17. Shoukry, M.M. Potentiometric studies of binary and ternary complexes of amoxycillin. Talanta 1992, 39, 1625–1628. [Google Scholar] [CrossRef]
  18. Zayed, M.A.; Abdallah, S.M. Synthesis and structure investigation of the antibiotic amoxicillin complexes of d-block elements. Spectrochim. Acta Part A 2005, 61, 2231–2238. [Google Scholar] [CrossRef]
  19. Cardiano, P.; Cigala, R.M.; Crea, F.; De Stefano, C.; Giuffrè, O.; Sammartano, S.; Vianelli, G. Potentiometric, UV and 1H NMR study on the interaction of penicillin derivatives with Zn(II) in aqueous solution. Biophys. Chem. 2017, 223, 1–10. [Google Scholar] [CrossRef] [PubMed]
  20. Cardiano, P.; Crea, F.; Foti, C.; Giuffrè, O.; Sammartano, S. Potentiometric, UV and 1H NMR study on the interaction of Cu2+ with ampicillin and amoxicillin in aqueous solution. Biophys. Chem. 2017, 224, 59–66. [Google Scholar] [CrossRef]
  21. Cigala, R.M.; Crea, F.; De Stefano, C.; Sammartano, S.; Vianelli, G. Thermodynamic Parameters for the Interaction of Amoxicillin andAmpicillin with Magnesium in NaCl Aqueous Solution, at Different Ionic Strengths and Temperatures. J. Chem. Eng. Data 2017, 62, 1018–1027. [Google Scholar] [CrossRef]
  22. Giuffrè, O.; Angowska, S.; Foti, C.; Sammartano, S. Thermodynamic Study on the Interaction of Ampicillin and Amoxicillin with Ca2+ in Aqueous Solution at Different Ionic Strengths and Temperatures. J. Chem. Eng. Data 2019, 64, 800–809. [Google Scholar] [CrossRef]
  23. Kupka, T. b-Lactam antibiotics. Spectroscopy and molecular orbital (MO) calculations: Part I: IR studies of complexation in penicillin-transition metal ion systems and semi-empirical PM3 calculations on simple model compounds. Spectrochim. Acta A 1997, 53, 2649–2658. [Google Scholar] [CrossRef]
  24. Filella, M.; May, P.M. Reflections on the calculation and publication of potentiometrically-determined formation constants. Talanta 2005, 65, 1221–1225. [Google Scholar] [CrossRef]
  25. Furia, E.; Napoli, A.; Tagarelli, A.; Sindona, G. Speciation of 2-hydroxybenzoic acid with Calcium(II), Magnesium(II), and Nickel(II) Cations in self-medium. J. Chem. Eng. Data 2013, 58, 1349–1353. [Google Scholar] [CrossRef]
  26. Cardiano, P.; Giacobello, F.; Giuffrè, O.; Sammartano, S. Thermodynamics of Al3+-thiocarboxylate interaction in aqueous solution. J. Mol. Liq. 2016, 222, 614–621. [Google Scholar] [CrossRef]
  27. De Stefano, C.; Foti, C.; Giuffrè, O.; Sammartano, S. Acid-base and UV behaviour of 3-(3,4-dihydroxyphenyl)-propenoic acid (caffeic acid) and complexing ability towards different divalent metal cations in aqueous solution. J. Mol. Liq. 2014, 195, 9–16. [Google Scholar] [CrossRef]
  28. Falcone, G.; Giuffrè, O.; Sammartano, S. Acid-base and UV properties of some aminophenol ligands and their complexing ability towards Zn2+ in aqueous solution. J. Mol. Liq. 2011, 159, 146–151. [Google Scholar] [CrossRef]
  29. Cardiano, P.; De Stefano, C.; Giuffrè, O.; Sammartano, S. Thermodynamic and spectroscopic study for the interaction of dimethyltin(IV) with L-cysteine in aqueous solution. Biophys. Chem. 2008, 133, 19–27. [Google Scholar] [CrossRef] [Green Version]
  30. Cardiano, P.; Falcone, G.; Foti, C.; Giuffrè, O.; Sammartano, S. Methylmercury(II)-sulphur containing ligand interactions:A potentiometric, calorimetric and 1H-NMR study in aqueous solution. New J. Chem. 2011, 35, 800–806. [Google Scholar] [CrossRef]
  31. Cardiano, P.; Giuffrè, O.; Pellerito, L.; Pettignano, A.; Sammartano, S.; Scopelliti, M. Thermodynamic and spectroscopic study of the binding of dimethyltin(IV) by citrate at 25 °C. Appl. Organomet. Chem. 2006, 20, 425–435. [Google Scholar] [CrossRef]
  32. Cesario, D.; Furia, E.; Mazzone, G.; Beneduci, A.; De Luca, G.; Sicilia, E. Complexation of Al3+ and Ni2+ by L -Ascorbic Acid: An Experimental and Theoretical Investigation. J. Phys. Chem. A 2017, 121, 9773–9781. [Google Scholar] [CrossRef]
  33. Furia, E.; Sindona, G. Interaction of iron(III) with 2-hydroxybenzoic acid in aqueous solutions. J. Chem. Eng. Data 2012, 57, 195–199. [Google Scholar] [CrossRef]
  34. Cardiano, P.; Giuffrè, O.; Napoli, A.; Sammartano, S. Potentiometric, 1H-NMR, ESI-MS investigation on dimethyltin(IV) cation-mercaptocarboxylate interaction in aqueous solution. New J. Chem. 2009, 33, 2286–2295. [Google Scholar] [CrossRef]
  35. Gonzales, C.; Jourad, E.; Spinelli, S.; Thomas, O. UV-Visible Spectrophotometry of Water and Waste Water; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar]
  36. Cardiano, P.; Giacobello, F.; Giuffrè, O.; Sammartano, S. Thermodynamic and spectroscopic study on Al3+-polycarboxylate interaction in aqueous solution. J. Mol. Liq. 2017, 232, 45–54. [Google Scholar] [CrossRef]
  37. Cardiano, P.; Foti, C.; Giuffrè, O. On the interaction of N-acetylcysteine with Pb2+, Zn2+, Cd2+ and Hg2+. J. Mol. Liq. 2016, 223, 360–367. [Google Scholar] [CrossRef]
  38. Cardiano, P.; Giacobello, F.; Giuffrè, O.; Sammartano, S. Thermodynamic and spectroscopic study of Al3+ interaction with glycine, L-cysteine and tranexamic acid in aqueous solution. Biophys Chem. 2017, 230, 10–19. [Google Scholar] [CrossRef]
  39. Crea, F.; Crea, P.; De Stefano, C.; Giuffrè, O.; Pettignano, A.; Sammartano, S. Thermodynamic Parameters for the Protonation of Poly(allylamine) in concentrated LiCl(aq) and NaCl(aq). J. Chem. Eng. Data 2004, 49, 658–663. [Google Scholar] [CrossRef]
  40. Cardiano, P.; Falcone, G.; Foti, C.; Giuffrè, O.; Napoli, A. Binding ability of glutathione towards alkyltin(IV) compounds in aqueous solution. J. Inorg. Biochem. 2013, 129, 84–93. [Google Scholar] [CrossRef]
  41. Gianguzza, A.; Giuffrè, O.; Piazzese, D.; Sammartano, S. Aqueous solution chemistry of alkyltin(IV) compounds for speciation studies in biological fluids and natural waters. Coord. Chem. Rev. 2012, 256, 222–239. [Google Scholar] [CrossRef]
  42. Cardiano, P.; De Stefano, C.; Foti, C.; Giacobello, F.; Giuffrè, O.; Sammartano, S. Sequestration of HEDPA, NTA and phosphonic NTA derivatives towards Al3+ in aqueous solution. J. Mol. Liq. 2018, 261, 96–106. [Google Scholar] [CrossRef]
  43. Aiello, D.; Cardiano, P.; Cigala, R.M.; Gans, P.; Giacobello, F.; Giuffrè, O.; Napoli, A.; Sammartano, S. Sequestering Ability of Oligophosphate Ligands toward Al3+ in Aqueous Solution. J. Chem. Eng. Data 2017, 62, 3981–3990. [Google Scholar] [CrossRef]
  44. De Stefano, C.; Foti, C.; Giuffrè, O.; Milea, D. Complexation of Hg2+, CH3Hg+, Sn2+, and (CH3)2Sn2+ with phosphonic NTA derivatives. New J. Chem. 2016, 40, 1443–1453. [Google Scholar] [CrossRef]
  45. Cardiano, P.; Foti, C.; Giacobello, F.; Giuffrè, O.; Sammartano, S. Study of Al3+ interaction with AMP, ADP and ATP in aqueous solution. Biophys Chem. 2018, 234, 42–50. [Google Scholar] [CrossRef] [PubMed]
  46. Martell, A.E.; Smith, R.M.; Motekaitis, R.J. Critically Selected Stability Constants of Metal Complexes; National Institute of Standard Technology: Garthersburg, MD, USA, 2004. [Google Scholar]
  47. May, P.M.; Murray, K. Database of chemical reactions designed to achieve thermodynamic consistency automatically. J. Chem. Eng. Data 2001, 46, 1035–1040. [Google Scholar] [CrossRef]
  48. Pettit, L.D.; Powell, K.J. IUPAC Stability Constants Database Academic Software; IUPAC: Yorks, UK, 2001. [Google Scholar]
  49. De Stefano, C.; Sammartano, S.; Mineo, P.; Rigano, C. Computer Tools for the Speciation of Natural Fluids. In Marine Chemistry-An Environmental Analytical Chemistry Approach; Gianguzza, A., Pelizzetti, E., Sammartano, S., Eds.; Kluwer Academic Publishers: Amsterdam, The Netherlands, 1997; pp. 71–83. [Google Scholar]
  50. Gans, P.; Sabatini, A.; Vacca, A. Investigation of equilibria in solution. Determination of equilibrium constants with the HYPERQUAD suite of programs. Talanta 1996, 43, 1739–1753. [Google Scholar] [CrossRef]
  51. Gans, P.; Sabatini, A.; Vacca, A. Determination of equilibrium constants from spectrophometric data obtained from solutions of known pH: THE PROGRAM pHab. Ann. Chim. (Rome) 1999, 89, 45–49. [Google Scholar]
  52. Alderighi, L.; Gans, P.; Ienco, A.; Peters, D.; Sabatini, A.; Vacca, A. Hyperquad simulation and speciation (HySS): A utility program for the investigation of equilibria involving soluble and partially soluble species. Coord. Chem. Rev. 1999, 184, 311–318. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are not available from the authors.
Figure 1. Ligands under study.
Figure 1. Ligands under study.
Molecules 25 03110 g001
Figure 2. Speciation diagram of Mn2+-Amp(L) system, at CM = 2 mmol L−1, CL = 4 mmol L−1, t = 25 °C, I = 0.15 mol L−1 (solid line), I = 1 mol L−1 (dashed line).
Figure 2. Speciation diagram of Mn2+-Amp(L) system, at CM = 2 mmol L−1, CL = 4 mmol L−1, t = 25 °C, I = 0.15 mol L−1 (solid line), I = 1 mol L−1 (dashed line).
Molecules 25 03110 g002
Figure 3. Experimental spectra at different pH on solutions containing Mn2+(M) and Amp(L) at CM = 0.05 mmol L−1 and CL = 0.05 mmol L−1, t = 25 °C, I = 0.15 mol L−1 in NaCl.
Figure 3. Experimental spectra at different pH on solutions containing Mn2+(M) and Amp(L) at CM = 0.05 mmol L−1 and CL = 0.05 mmol L−1, t = 25 °C, I = 0.15 mol L−1 in NaCl.
Molecules 25 03110 g003
Figure 4. Molar absorbance values of protonated and unprotonated Amp species and Mn2+-Amp complexes at t = 25 °C and I = 0.15 mol L−1 in NaCl (charges omitted for simplicity).
Figure 4. Molar absorbance values of protonated and unprotonated Amp species and Mn2+-Amp complexes at t = 25 °C and I = 0.15 mol L−1 in NaCl (charges omitted for simplicity).
Molecules 25 03110 g004
Figure 5. Speciation diagram of Mn2+-Amox(L) system, at CM = 2 mmol L−1, CL = 4 mmol L−1, t = 25 °C, I = 0.15 mol L−1 (solid line), I = 1 mol L−1 (dashed line).
Figure 5. Speciation diagram of Mn2+-Amox(L) system, at CM = 2 mmol L−1, CL = 4 mmol L−1, t = 25 °C, I = 0.15 mol L−1 (solid line), I = 1 mol L−1 (dashed line).
Molecules 25 03110 g005
Figure 6. Experimental spectra at different pH on solutions containing Mn2+(M) and Amox(L) at CM = 0.05 mmol L−1 and CL = 0.075 mmol L−1, t = 25 °C, I = 0.15 mol L−1 in NaCl.
Figure 6. Experimental spectra at different pH on solutions containing Mn2+(M) and Amox(L) at CM = 0.05 mmol L−1 and CL = 0.075 mmol L−1, t = 25 °C, I = 0.15 mol L−1 in NaCl.
Molecules 25 03110 g006
Figure 7. Molar absorbance values of Mn2+-Amox complexes and protonated and unprotonated Amox species at t = 25 °C and I = 0.15 mol L−1 in NaCl (charges omitted for simplicity).
Figure 7. Molar absorbance values of Mn2+-Amox complexes and protonated and unprotonated Amox species at t = 25 °C and I = 0.15 mol L−1 in NaCl (charges omitted for simplicity).
Molecules 25 03110 g007
Figure 8. Bar plot of ΔG, ΔH, TΔS referring to the Mn2+-Amp and Mn2+-Amox systems at t = 25 °C, I = 0.15 mol L−1 in NaCl, according to the reaction (2).
Figure 8. Bar plot of ΔG, ΔH, TΔS referring to the Mn2+-Amp and Mn2+-Amox systems at t = 25 °C, I = 0.15 mol L−1 in NaCl, according to the reaction (2).
Molecules 25 03110 g008
Figure 9. Speciation diagram of Mn2+-Amp(L) system, at CM = 2 mmol L−1, CL = 4 mmol L−1, I = 0.15 mol L−1, t = 15 °C (solid line), t = 37 °C (dashed line).
Figure 9. Speciation diagram of Mn2+-Amp(L) system, at CM = 2 mmol L−1, CL = 4 mmol L−1, I = 0.15 mol L−1, t = 15 °C (solid line), t = 37 °C (dashed line).
Molecules 25 03110 g009
Figure 10. Speciation diagram of Mn2+-Amox(L) system, at CM = 2 mmol L−1, CL = 4 mmol L−1, I = 0.15 mol L−1, t = 15 °C (solid line), t = 37 °C (dashed line).
Figure 10. Speciation diagram of Mn2+-Amox(L) system, at CM = 2 mmol L−1, CL = 4 mmol L−1, I = 0.15 mol L−1, t = 15 °C (solid line), t = 37 °C (dashed line).
Molecules 25 03110 g010
Figure 11. Sum of the fractions of Mn2+-Amp (in red), Mn2+-Amox (in black) species at pH = 7.4, t = 37 °C, I = 0.15 mol L−1 in NaCl.
Figure 11. Sum of the fractions of Mn2+-Amp (in red), Mn2+-Amox (in black) species at pH = 7.4, t = 37 °C, I = 0.15 mol L−1 in NaCl.
Molecules 25 03110 g011
Table 1. Experimental formation constant values for Mn2+-Amp(L), Mn2+-Amox(L) species obtained by potentiometry at different temperatures and ionic strengths.
Table 1. Experimental formation constant values for Mn2+-Amp(L), Mn2+-Amox(L) species obtained by potentiometry at different temperatures and ionic strengths.
LSpecies logβ a
t = 15 °C t = 25 °C t = 37 °C
I = 0.15 bI = 0.15 bI = 0.48 bI = 0.96 bI = 0.15 b
AmpMnLH9.47(2) c9.47(1) c9.57(3) c8.68(8) c8.18(8) c
MnL2.36(3)2.32(2)2.37(5)2.14(3)2.09(3)
AmoxMnLH219.92(2)19.64(2)19.48(4)20.14(2)19.55(2)
MnLH12.25(3)12.76(1)12.80(2)12.44(4)12.59(2)
logK d
AmpMnLH2.102.422.471.441.42
MnL2.362.322.372.142.09
AmoxMnLH22.632.782.673.173.16
MnLH2.543.203.322.913.19
a Refer to the reaction (1); b in mol L−1; c ≥95% of confidence interval; d refer to the reaction (2).
Table 2. Comparison between experimental formation constant values of Mn2+-Amp species obtained by spectrophotometry and potentiometry at t = 25 °C and I = 0.15 mol L−1.
Table 2. Comparison between experimental formation constant values of Mn2+-Amp species obtained by spectrophotometry and potentiometry at t = 25 °C and I = 0.15 mol L−1.
LSpecieslogβ a
SpectrophotometryPotentiometry
AmpMLH9.53(2) b9.47(2) b
ML2.30(2)2.32(2)
AmoxMLH219.77(7)19.64(2)
MLH12.96(3)12.76(1)
a Refer to the reaction (1); b ≥95% of confidence interval.
Table 3. Thermodynamic formation parameters for Mn2+-Amp and Mn2+-Amox species.
Table 3. Thermodynamic formation parameters for Mn2+-Amp and Mn2+-Amox species.
SpeciesLlogβ0 aCb
MLHAmp10.0(1) c−0.8(2) c
ML2.88(4)0.10(9)
MLH2Amox20.13(8)1.2(1)
MLH13.8(1)0.4(2)
a Refers to reaction (1); b Equation (3); c ≥95% of confidence interval.
Table 4. Thermodynamic formation parameters for Mn2+-Amp, Mn2+-Amox species at t = 25 °C and I = 0.15 mol L−1.
Table 4. Thermodynamic formation parameters for Mn2+-Amp, Mn2+-Amox species at t = 25 °C and I = 0.15 mol L−1.
SpeciesL−ΔG a,bΔH a,bTΔSa,b
MLHAmp54.1−101(13) c−47
ML13.2−21(3)−8
MLH2Amox112.1−28(6)84
MLH72.820(5)93
ReactionL−ΔG bΔH bTΔSb
M + LHAmp13.8−57−43
M + L13.2−21−8
M + LH2Amox15.94359
M + LH18.3−513
a Refer to the reaction (1); b Expressed in kJ mol−1; c ≥95% of confidence interval.
Table 5. pL0.5 values of Amp and Amox towards Mn2+ at different pH, ionic strengths and temperatures.
Table 5. pL0.5 values of Amp and Amox towards Mn2+ at different pH, ionic strengths and temperatures.
Lt/°CI/mol L−1pHpL0.5
Amp250.157.42.35
250.57.42.41
2517.42.17
150.157.42.28
370.157.42.01
370.158.02.04
370.156.51.78
Amox250.157.43.05
250.57.43.13
2517.43.06
150.157.42.60
370.157.43.17
370.158.03.14
370.156.53.16

Share and Cite

MDPI and ACS Style

Foti, C.; Giuffrè, O. Interaction of Ampicillin and Amoxicillin with Mn2+: A Speciation Study in Aqueous Solution. Molecules 2020, 25, 3110. https://doi.org/10.3390/molecules25143110

AMA Style

Foti C, Giuffrè O. Interaction of Ampicillin and Amoxicillin with Mn2+: A Speciation Study in Aqueous Solution. Molecules. 2020; 25(14):3110. https://doi.org/10.3390/molecules25143110

Chicago/Turabian Style

Foti, Claudia, and Ottavia Giuffrè. 2020. "Interaction of Ampicillin and Amoxicillin with Mn2+: A Speciation Study in Aqueous Solution" Molecules 25, no. 14: 3110. https://doi.org/10.3390/molecules25143110

Article Metrics

Back to TopTop