Marija Bester-Rogac | University of Ljubljana (original) (raw)
Papers by Marija Bester-Rogac
International Journal of Pharmaceutics, 2006
Small-angle X-ray scattering technique has been used to study the structural properties of the qu... more Small-angle X-ray scattering technique has been used to study the structural properties of the quaternary microemulsion Tween 40 ® /Imwitor 308 ® /isopropyl myristate/water and of five-component system obtained by the addition of the drug ketoprofen to the original quaternary system. The results enlighten the structuration of the studied systems and represent new complementary findings to the previous study [Podlogar, F., Bešter-Rogač, M., Gašperlin, M., 2005. The effect of internal structure of selected water-Tween 40 (R)-Imwitor 308 (R)-IPM microemulsions on ketoprofene release. Int. J. Pharm. 302, 68-77] on the correlation between the structuration of these systems and the release rates of the ketoprofen. The present results indicate that in the samples with the moderate to high concentration of water where the latter is a continuous phase the addition of smaller amounts of the ketoprofen does not change their inner structuration significantly. The quaternary sample containing 46.2 wt.% of water seems to be very near the composition where the transition from the bicontinuous to the lamellar structure of the microemulsion occurs. In the samples containing from 46.2 to 62.7 wt.% of water the swelling of lamellar phases with constant thickness of double-layer can be characterized. At approximately the latter composition another noticeable transition in the inner structuration of the microemulsion has been observed. Interestingly, all these changes in the inner structuration of the studied systems did not affect the trend of the drug release rates in this regime of water concentrations.
The Journal of Chemical Thermodynamics, 2015
The micellization behaviour of the long-chain carboxylates: sodium octanoate (NaC8), sodium decan... more The micellization behaviour of the long-chain carboxylates: sodium octanoate (NaC8), sodium decanoate (NaC10), potassium decanoate (KC10), caesium decanoate (CsC10) and sodium dodecanoate (NaC12) in aqueous solutions were studied using electrical conductivity measurements between the temperatures (278.15 and 328.15) K. By using the pseudo-phase separation model and Gibbs-Helmholtz equation, the thermodynamic parameters for the micellization process were obtained from the temperature dependence of the critical micelle concentration, CMC, and the degree of micelle ionisation, b. The heat capacity of micellization, D mic c o p , estimated from the temperature dependence of enthalpy of micellization, D mic H o , were correlated with the size of the non-polar accessible surface area of the solvent, which is removed from contact with water during micellization. In all systems investigated, results suggest that the micelle core is still in contact with water molecules.
International Journal of Pharmaceutics, 2004
Pharmaceutically usable microemulsion systems were prepared from water and isopropyl myristate wi... more Pharmaceutically usable microemulsion systems were prepared from water and isopropyl myristate with a constant amount of Tween 40 and Imwitor 308 at a mass ratio of 1. Their type and structure were examined by measuring density and surface tension, and by viscometry, electric conductivity, differential scanning calorimetry (DSC) and small-angle X-ray scattering (SAXS), and the degree of agreement between the techniques was assessed. A model based on monodisperse hard spheres adequately fits the SAXS data in W/O microemulsions predicting, depending on composition, elongated or spherical droplets. It also suggests the involvement of strong attractive interactions in O/W systems. Results of conductivity, viscosity, density and surface tension measurements confirm the prediction of a percolation transition to a bicontinuous structure. DSC detects the degree of water interaction with surfactants thus identifying the type of microemulsion. The conclusions from all the techniques agree well and indicate that such studies could also be carried out on more complex systems. In future, the ability to determine type and structure of such microemulsion systems could enable partitioning and release rates of drugs from microemulsions to be predicted.
International Journal of Pharmaceutics, 2005
Microemulsions are a promising vehicle for administrating drugs. In order to lay the basis for pr... more Microemulsions are a promising vehicle for administrating drugs. In order to lay the basis for predicting drug release under in vivo conditions, where the microemulsion composition is continuously varying, we have studied the release of ketoprofene as a model drug, from microemulsions on a dilution line containing, initially, 20 wt.% of isopropyl miristate (IPM) and 80 wt.% of the surfactant (Tween 40 ® )/co-surfactant (Imwitor 308 ® ), 1:1 wt.% mixture. Mixture compositions corresponding to the different types and structure of microemulsion were identified by measuring density, surface tension, electric conductivity, pH and differential scanning calorimetry. Ketoprofene release was then measured for each type and structure. The main factor influencing ketoprofene release was shown to be the strength of the interactions between microemulsion components. Strong interactions prevented rapid ketoprofene release in the water-in oil region, although the release was not dependent on the degree of percolation. Release kinetics in all cases follow zero order kinetics, indicating that the release rate is dependent on the diffusion of ketoprofene inside the microemulsion carrier. Combining different methods to obtain the physical and structural properties of microemulsions can be thus used to predict the release of ketoprofen from a microemulsion. (M. Gašperlin).
The Journal of Chemical Thermodynamics, 2011
The thermodynamics of micelle formation of the cationic surfactants decyltrimethylammonium chlori... more The thermodynamics of micelle formation of the cationic surfactants decyltrimethylammonium chloride (DeTAC) and tetradecyltrimethylammonium chloride (TTAC) in water and aqueous NaCl solutions have been investigated. Isothermal titration calorimetry (ITC) was used to study the effect of added salt on the critical micelle concentration, cmc, and enthalpy of micellization, ΔmicH∘, between the temperatures (278.15 and 328.15)K. From the temperature dependence of
The Journal of Physical Chemistry B, 2014
Molar conductivities, Λ, of dilute solutions of the ionic liquids (ILs) 1-ethyl-3-methylimidazoli... more Molar conductivities, Λ, of dilute solutions of the ionic liquids (ILs) 1-ethyl-3-methylimidazolium tetrafluoroborate ([emim][BF 4 ]), 1-butyl-3-methylimidazolium tetrafluoroborate ([bmim][BF 4 ]), 1-butyl-3-methylimidazolium hexafluorophosphate ([bmim][PF 6 ]), 1-hexyl-3-methylimidazolium tetrafluoroborate ([hmim][BF 4 ]), and 1-hexyl-3-methylimidazolium bis-(trifluoromethanesulfonyl)amide ([hmim][NTf 2 ]) in acetonitrile were determined as a function of temperature in the range 273.15−313.15 K. The data were analyzed with Barthel's lcCM model to obtain limiting molar conductivities, Λ ∞ (T), and association constants, K A°( T) of these electrolytes. The temperature dependence of these parameters, as well as the extracted limiting cation conductivities, λ i ∞ , were discussed. Additionally, dielectric spectra for [hmim][NTf 2 ] + AN were analyzed in terms of ion association and ion solvation and compared with the inference from conductivity. It appears that in dilute solutions the imidazolium ring of the cations is solvated by ∼6 AN molecules that are slowed by a factor of ∼8−10 compared to the bulk-solvent dynamics. Ion association of imidazolium ILs to contact ion pairs is only moderate, similar to common 1:1 electrolytes in this solvent.
The Journal of Physical Chemistry B, 2009
The conductance of poly(anetholesulfonic acid) and its lithium, sodium, and cesium salts in water... more The conductance of poly(anetholesulfonic acid) and its lithium, sodium, and cesium salts in water was measured in the range from c(m) to approximately 0.001 to 0.20 monomol/dm3 and in the temperature range from T = 278 to 308 K. For the alkaline salts of poly(anetholesulfonic acid) Walden's rule is satisfied reasonably well, but not for the polyacid itself. For the sodium salt of poly(anetholesulfonic acid) we determined the concentration dependence of the polyion transference number at 298 K. From the measurements we calculated the fraction of free sodium ions, alpha, in the solution. The results were analyzed theoretically, using the expression alpha approximately D+/D0+, where D+ is the self-diffusion coefficient of the counterion species. The cylindrical cell model and the Poisson-Boltzmann theory were used to calculate D+/D0+. The calculations are in good qualitative agreement with experimental data. These new measurements for polyanetholesulfonates were compared with the experimental results for poly(styrenesulfonic acid) and its salts obtained from the literature. The conductivities of aqueous solutions of poly(anetholesulfonic acid) and its salt are higher than the corresponding polystyrenesulfonate solutions. This can be explained by a smaller fraction of "free" (conducting) counterions in the latter case. This finding is consistent with thermodynamic data for these solutions as well as with the transference number measurements for sodium polyanetholesulfonate solutions presented here.
The Journal of Physical Chemistry B, 2004
ABSTRACT
The Journal of Physical Chemistry B, 2012
Precise measurements of electrical conductivities of methylparaben, ethylparaben, propylparaben, ... more Precise measurements of electrical conductivities of methylparaben, ethylparaben, propylparaben, and butylparaben sodium salts in dilute aqueous solutions were performed from 278.15 to 313.15 K in 5 K intervals. Experimental conductivity data were analyzed applying the Quint-Viallard conductivity equations by taking into account the salt hydrolysis in aqueous solutions. These evaluations yield the limiting conductances of paraben anions and the dissociation constants of the investigated parabens in water. From temperature dependence of dissociation constants, the thermodynamic functions associated with the dissociation process were estimated. It was discovered that the contributions of enthalpy and entropy to the Gibbs free energy are quite similar. The Walden products of paraben anions in water are independent of temperature, indicating that the hydrodynamic radii are not significantly affected by temperature.
The Journal of Physical Chemistry B, 2013
Systematic and precise measurements of electrical conductivities of aqueous solutions of cadmium ... more Systematic and precise measurements of electrical conductivities of aqueous solutions of cadmium chloride were performed in the 2 × 10(-5)-1 × 10(-2) mol·dm(-3) concentration range, from 278.15 to 313.15 K. Determined conductances were interpreted in terms of molecular model which includes a mixture of two 1:1 and 2:1 electrolytes. The molar limiting conductances of λ(0)(CdCl(+), T) and λ(0)(1/2Cd(2+), T), the equilibrium constants of CdCl(+) formation K(T) and the corresponding standard thermodynamic functions were evaluated using the Quint-Viallard conductivity equations, the Debye-Hückel equations for activity coefficients and the mass-action equation. An excellent agreement between calculated and experimental conductivities was reached.
The Journal of Physical Chemistry B, 2006
A general approach is proposed to analyze electrical conductivities in aqueous solutions of polyb... more A general approach is proposed to analyze electrical conductivities in aqueous solutions of polybasic organic acids. Experimental conductivities are examined in the context of dissociation and hydrolysis reactions by applying the Quint-Viallard conductivity equations and the Debye-Hückel equations for activity coefficients. The proposed numerical procedure is illustrated by the case of benzenehexacarboxylic (mellitic) acid and its neutral and acidic salts. From conductivity measurements of mellitic acid and its salts, performed in dilute aqueous solutions in the 278.15-308.15 K temperature range, the limiting conductances of mellitic anions, lambda(0)(1/jH(6-j)Mel(-j), T), j = 1, 2, 3, 4, 5, 6 are determined.
The Journal of Physical Chemistry B, 2007
Systematic determinations of electrical conductivities of sodium penicillin G, potassium penicill... more Systematic determinations of electrical conductivities of sodium penicillin G, potassium penicillin G, and potassium penicillin V in the 278.15-313.15 K temperature range are reported. These conductivities are examined by applying the Quint-Viallard conductivity equations and the Debye-Hückel equations for activity coefficients. Determined dissociation constants and the limiting conductances of penicillin anions are based on the assumption that in dilute aqueous solutions, penicillin salts behave as acidic salts of dibasic acids, which are the final products of degradation reactions in acidic media. conductivity equations, 39-42 taking into account the dissociation of likely products of penicillin hydrolysis and degradation reactions.
PLoS ONE, 2012
Background: ATP-dependent D-alanine:D-alanine ligase (Ddl) is a part of biochemical machinery inv... more Background: ATP-dependent D-alanine:D-alanine ligase (Ddl) is a part of biochemical machinery involved in peptidoglycan biosynthesis, as it catalyzes the formation of the terminal D-ala-D-ala dipeptide of the peptidoglycan precursor UDPMurNAcpentapeptide. Inhibition of Ddl prevents bacterial growth, which makes this enzyme an attractive and viable target in the urgent search of novel effective antimicrobial drugs. To address the problem of a relentless increase in resistance to known antimicrobial agents we focused our attention to discovery of novel ATP-competitive inhibitors of Ddl.
Monatshefte für Chemie - Chemical Monthly, 2011
Abstract The density, refractive index, and electrical permittivity of 1,4-dioxane solutions of ... more Abstract The density, refractive index, and electrical permittivity of 1,4-dioxane solutions of acesulfame [6-methyl-1,2,3-oxathiazine-4(3H)-one-2,2-dioxide] and saccharin [1,2-benzisothiazole-3(2H)-one-1,1-dioxide] were measured at 298.15 K. From the experimental data the limiting apparent specific volume, refraction, and polarization of acesulfame and saccharin were calculated. The electrical dipole moments of acesulfame and saccharin were estimated according to the Debye, Onsager, and Kirkwood approaches. The association via dipole–dipole
Langmuir, 2012
A systematic investigation of the micellization process of a biocompatible zwitterionic surfactan... more A systematic investigation of the micellization process of a biocompatible zwitterionic surfactant 3-[(3-cholamidopropyl)dimethylammonium]-1-propanesulfonate (CHAPS) has been carried out by isothermal titration calorimetry (ITC) at temperatures between 278.15 K and 328.15 K in water, aqueous NaCl (0.1, 0.5, and 1 M), and buffer solutions (pH = 3.0, 6.8, and 7.8). The effect of different cations and anions on the micellization of CHAPS surfactant has been also examined in LiCl, CsCl, NaBr, and NaI solutions at 308.15 K. It turned out that the critical micelle concentration, cmc, is only slightly shifted toward lower values in salt solutions, whereas in buffer media it remains similar to its value in water. From the results obtained, it could be assumed that CHAPS behaves as a weakly charged cationic surfactant in salt solutions and as a nonionic surfactant in water and buffer medium. Conventional surfactants alike, CHAPS micellization is endothermic at low and exothermic at high temperatures, but the estimated enthalpy of micellization, ΔH M 0 , is considerably lower in comparison with that obtained for ionic surfactants in water and NaCl solutions. The standard Gibbs free energy, ΔG M 0 , and entropy, ΔS M 0 , of micellization were estimated by fitting the model equation based on the mass action model to the experimental data. The aggregation numbers of CHAPS surfactant around cmc, obtained by the fitting procedure also, are considerably low (n agg ≈ 5 ± 1). Furthermore, some predictions about the hydration of the micelle interior based on the correlation between heat capacity change, Δc p,M 0 , and changes in solvent-accessible surface upon micelle formation were made. CHAPS molecules are believed to stay in contact with water upon aggregation, which is somehow similar to the micellization process of short alkyl chain cationic surfactants.
Langmuir, 2013
Specific effects of the sodium salts of m- and p-hydroxybenzoates (m-HB and p-HB) on the aggregat... more Specific effects of the sodium salts of m- and p-hydroxybenzoates (m-HB and p-HB) on the aggregation process of dodecyltrimethylammonium chloride have been investigated by isothermal titration calorimetry, electrical conductivity, and (1)H NMR and compared with already reported data for the sodium salt of o-hydroxybenzoate (o-HB). For p-HB, it has been found that the aggregate is only formed by spherical micelles at all p-HB concentrations. On the other side, the situation is more complex for o-HB, where two distinct states of aggregation can be involved, depending on the concentration of o-HB. At high salt concentration, rodlike micelles are formed, whereas at lower concentration spherical aggregates are predominant. The transition from the cylinder to the sphere increases the mobility of the surfactant because the core of the rodlike micelles is more closely packed due to the expulsion of water from the interior of the aggregate. m-HB exhibits an intermediate behavior between these two extreme situations. The effect of the position of hydrophilic substituents on the aromatic ring on the insertion of the hydroxybenzoate anion in the micellar aggregate has been discussed.
Journal of Solution Chemistry, 2008
The electrical conductivity of aqueous cyclamic acid was studied in the concentration range (0.00... more The electrical conductivity of aqueous cyclamic acid was studied in the concentration range (0.0004 < c/mol·dm −3 < 0.14) at temperatures ranging from 278.15 to 303.15 K. Conductivities were measured by a precise method and examined by applying extended conductivity equations taking into account dimerization and incomplete electrolyte dissociation. Limiting molar electrolyte and ionic conductivities, dissociation and dimerization constants, and thermodynamic functions associated with dissociation and dimerization of cyclamic acid are discussed.
Journal of Solution Chemistry, 2008
The viscosities of aqueous solutions of lithium, sodium, potassium, rubidium and caesium cyclohex... more The viscosities of aqueous solutions of lithium, sodium, potassium, rubidium and caesium cyclohexylsulfamates were measured at 293.15, 298.15, 303.15, 313.15 and 323.15 K. The relative viscosity data were analyzed and interpreted in terms of the Kaminsky equation, η r = 1 + Ac 1/2 + Bc + Dc 2 . The viscosity A-coefficient was calculated from the Falkenhagen-Dole theory. The viscosity B-coefficients are positive and relatively large. Their temperature coefficient ∂B/∂T is negative or near zero for lithium and sodium salts whereas for potassium, rubidium and caesium salts it is positive. The viscosity D-coefficient is positive. This was explained by the size of the ions, structural solute-solute interactions, hydrodynamic effect, and by higher terms of the long-range Debye-Hückel type of forces. From the viscosity B-coefficients the thermodynamic functions of activation of viscous flow were calculated. The limiting partial molar Gibbs energy of activation of viscous flow of the solute was divided into contributions due to solvent molecules and the solute in the transition state. The activation energy of the solvent molecules was calculated using the limiting Gibbs energy of activation for the conductance of the solute ions. The activation energy of the solvent molecules was then discussed in terms of the nature of the alkali-metal ions and their influence on the structure of water. The limiting activation entropy and enthalpy of the solute for activation of viscous flow were interpreted by ion-solvent bond formation or breaking in the transition state of the solvent. The hydration numbers of the investigated electrolytes were calculated from the specific viscosity of the solutions.
Journal of Solution Chemistry, 2005
The electric conductivities of aqueous solutions of the lithium, sodium, potassium and ammonium s... more The electric conductivities of aqueous solutions of the lithium, sodium, potassium and ammonium salts of cyclohexylsulfamic acid were measured from 5 to 35 • C (in steps of 5 • C) in the concentration range 3 × 10 −4 < c/mol-dm −3 < 0.01. Data analysis based on a chemical model of electrolyte solutions yielded the limiting molar conductance ∞ and the association constant K A . Using the known values of the limiting conductances of lithium, sodium and potassium ions, the limiting conductances of the cyclohexylsulfamic ion were evaluated. Total dissociation of the investigated salts in water and negligible hydration of the cyclohexylsulfamate anion are evident.
International Journal of Pharmaceutics, 2006
Small-angle X-ray scattering technique has been used to study the structural properties of the qu... more Small-angle X-ray scattering technique has been used to study the structural properties of the quaternary microemulsion Tween 40 ® /Imwitor 308 ® /isopropyl myristate/water and of five-component system obtained by the addition of the drug ketoprofen to the original quaternary system. The results enlighten the structuration of the studied systems and represent new complementary findings to the previous study [Podlogar, F., Bešter-Rogač, M., Gašperlin, M., 2005. The effect of internal structure of selected water-Tween 40 (R)-Imwitor 308 (R)-IPM microemulsions on ketoprofene release. Int. J. Pharm. 302, 68-77] on the correlation between the structuration of these systems and the release rates of the ketoprofen. The present results indicate that in the samples with the moderate to high concentration of water where the latter is a continuous phase the addition of smaller amounts of the ketoprofen does not change their inner structuration significantly. The quaternary sample containing 46.2 wt.% of water seems to be very near the composition where the transition from the bicontinuous to the lamellar structure of the microemulsion occurs. In the samples containing from 46.2 to 62.7 wt.% of water the swelling of lamellar phases with constant thickness of double-layer can be characterized. At approximately the latter composition another noticeable transition in the inner structuration of the microemulsion has been observed. Interestingly, all these changes in the inner structuration of the studied systems did not affect the trend of the drug release rates in this regime of water concentrations.
The Journal of Chemical Thermodynamics, 2015
The micellization behaviour of the long-chain carboxylates: sodium octanoate (NaC8), sodium decan... more The micellization behaviour of the long-chain carboxylates: sodium octanoate (NaC8), sodium decanoate (NaC10), potassium decanoate (KC10), caesium decanoate (CsC10) and sodium dodecanoate (NaC12) in aqueous solutions were studied using electrical conductivity measurements between the temperatures (278.15 and 328.15) K. By using the pseudo-phase separation model and Gibbs-Helmholtz equation, the thermodynamic parameters for the micellization process were obtained from the temperature dependence of the critical micelle concentration, CMC, and the degree of micelle ionisation, b. The heat capacity of micellization, D mic c o p , estimated from the temperature dependence of enthalpy of micellization, D mic H o , were correlated with the size of the non-polar accessible surface area of the solvent, which is removed from contact with water during micellization. In all systems investigated, results suggest that the micelle core is still in contact with water molecules.
International Journal of Pharmaceutics, 2004
Pharmaceutically usable microemulsion systems were prepared from water and isopropyl myristate wi... more Pharmaceutically usable microemulsion systems were prepared from water and isopropyl myristate with a constant amount of Tween 40 and Imwitor 308 at a mass ratio of 1. Their type and structure were examined by measuring density and surface tension, and by viscometry, electric conductivity, differential scanning calorimetry (DSC) and small-angle X-ray scattering (SAXS), and the degree of agreement between the techniques was assessed. A model based on monodisperse hard spheres adequately fits the SAXS data in W/O microemulsions predicting, depending on composition, elongated or spherical droplets. It also suggests the involvement of strong attractive interactions in O/W systems. Results of conductivity, viscosity, density and surface tension measurements confirm the prediction of a percolation transition to a bicontinuous structure. DSC detects the degree of water interaction with surfactants thus identifying the type of microemulsion. The conclusions from all the techniques agree well and indicate that such studies could also be carried out on more complex systems. In future, the ability to determine type and structure of such microemulsion systems could enable partitioning and release rates of drugs from microemulsions to be predicted.
International Journal of Pharmaceutics, 2005
Microemulsions are a promising vehicle for administrating drugs. In order to lay the basis for pr... more Microemulsions are a promising vehicle for administrating drugs. In order to lay the basis for predicting drug release under in vivo conditions, where the microemulsion composition is continuously varying, we have studied the release of ketoprofene as a model drug, from microemulsions on a dilution line containing, initially, 20 wt.% of isopropyl miristate (IPM) and 80 wt.% of the surfactant (Tween 40 ® )/co-surfactant (Imwitor 308 ® ), 1:1 wt.% mixture. Mixture compositions corresponding to the different types and structure of microemulsion were identified by measuring density, surface tension, electric conductivity, pH and differential scanning calorimetry. Ketoprofene release was then measured for each type and structure. The main factor influencing ketoprofene release was shown to be the strength of the interactions between microemulsion components. Strong interactions prevented rapid ketoprofene release in the water-in oil region, although the release was not dependent on the degree of percolation. Release kinetics in all cases follow zero order kinetics, indicating that the release rate is dependent on the diffusion of ketoprofene inside the microemulsion carrier. Combining different methods to obtain the physical and structural properties of microemulsions can be thus used to predict the release of ketoprofen from a microemulsion. (M. Gašperlin).
The Journal of Chemical Thermodynamics, 2011
The thermodynamics of micelle formation of the cationic surfactants decyltrimethylammonium chlori... more The thermodynamics of micelle formation of the cationic surfactants decyltrimethylammonium chloride (DeTAC) and tetradecyltrimethylammonium chloride (TTAC) in water and aqueous NaCl solutions have been investigated. Isothermal titration calorimetry (ITC) was used to study the effect of added salt on the critical micelle concentration, cmc, and enthalpy of micellization, ΔmicH∘, between the temperatures (278.15 and 328.15)K. From the temperature dependence of
The Journal of Physical Chemistry B, 2014
Molar conductivities, Λ, of dilute solutions of the ionic liquids (ILs) 1-ethyl-3-methylimidazoli... more Molar conductivities, Λ, of dilute solutions of the ionic liquids (ILs) 1-ethyl-3-methylimidazolium tetrafluoroborate ([emim][BF 4 ]), 1-butyl-3-methylimidazolium tetrafluoroborate ([bmim][BF 4 ]), 1-butyl-3-methylimidazolium hexafluorophosphate ([bmim][PF 6 ]), 1-hexyl-3-methylimidazolium tetrafluoroborate ([hmim][BF 4 ]), and 1-hexyl-3-methylimidazolium bis-(trifluoromethanesulfonyl)amide ([hmim][NTf 2 ]) in acetonitrile were determined as a function of temperature in the range 273.15−313.15 K. The data were analyzed with Barthel's lcCM model to obtain limiting molar conductivities, Λ ∞ (T), and association constants, K A°( T) of these electrolytes. The temperature dependence of these parameters, as well as the extracted limiting cation conductivities, λ i ∞ , were discussed. Additionally, dielectric spectra for [hmim][NTf 2 ] + AN were analyzed in terms of ion association and ion solvation and compared with the inference from conductivity. It appears that in dilute solutions the imidazolium ring of the cations is solvated by ∼6 AN molecules that are slowed by a factor of ∼8−10 compared to the bulk-solvent dynamics. Ion association of imidazolium ILs to contact ion pairs is only moderate, similar to common 1:1 electrolytes in this solvent.
The Journal of Physical Chemistry B, 2009
The conductance of poly(anetholesulfonic acid) and its lithium, sodium, and cesium salts in water... more The conductance of poly(anetholesulfonic acid) and its lithium, sodium, and cesium salts in water was measured in the range from c(m) to approximately 0.001 to 0.20 monomol/dm3 and in the temperature range from T = 278 to 308 K. For the alkaline salts of poly(anetholesulfonic acid) Walden&amp;amp;amp;amp;amp;#39;s rule is satisfied reasonably well, but not for the polyacid itself. For the sodium salt of poly(anetholesulfonic acid) we determined the concentration dependence of the polyion transference number at 298 K. From the measurements we calculated the fraction of free sodium ions, alpha, in the solution. The results were analyzed theoretically, using the expression alpha approximately D+/D0+, where D+ is the self-diffusion coefficient of the counterion species. The cylindrical cell model and the Poisson-Boltzmann theory were used to calculate D+/D0+. The calculations are in good qualitative agreement with experimental data. These new measurements for polyanetholesulfonates were compared with the experimental results for poly(styrenesulfonic acid) and its salts obtained from the literature. The conductivities of aqueous solutions of poly(anetholesulfonic acid) and its salt are higher than the corresponding polystyrenesulfonate solutions. This can be explained by a smaller fraction of &amp;amp;amp;amp;amp;quot;free&amp;amp;amp;amp;amp;quot; (conducting) counterions in the latter case. This finding is consistent with thermodynamic data for these solutions as well as with the transference number measurements for sodium polyanetholesulfonate solutions presented here.
The Journal of Physical Chemistry B, 2004
ABSTRACT
The Journal of Physical Chemistry B, 2012
Precise measurements of electrical conductivities of methylparaben, ethylparaben, propylparaben, ... more Precise measurements of electrical conductivities of methylparaben, ethylparaben, propylparaben, and butylparaben sodium salts in dilute aqueous solutions were performed from 278.15 to 313.15 K in 5 K intervals. Experimental conductivity data were analyzed applying the Quint-Viallard conductivity equations by taking into account the salt hydrolysis in aqueous solutions. These evaluations yield the limiting conductances of paraben anions and the dissociation constants of the investigated parabens in water. From temperature dependence of dissociation constants, the thermodynamic functions associated with the dissociation process were estimated. It was discovered that the contributions of enthalpy and entropy to the Gibbs free energy are quite similar. The Walden products of paraben anions in water are independent of temperature, indicating that the hydrodynamic radii are not significantly affected by temperature.
The Journal of Physical Chemistry B, 2013
Systematic and precise measurements of electrical conductivities of aqueous solutions of cadmium ... more Systematic and precise measurements of electrical conductivities of aqueous solutions of cadmium chloride were performed in the 2 × 10(-5)-1 × 10(-2) mol·dm(-3) concentration range, from 278.15 to 313.15 K. Determined conductances were interpreted in terms of molecular model which includes a mixture of two 1:1 and 2:1 electrolytes. The molar limiting conductances of λ(0)(CdCl(+), T) and λ(0)(1/2Cd(2+), T), the equilibrium constants of CdCl(+) formation K(T) and the corresponding standard thermodynamic functions were evaluated using the Quint-Viallard conductivity equations, the Debye-Hückel equations for activity coefficients and the mass-action equation. An excellent agreement between calculated and experimental conductivities was reached.
The Journal of Physical Chemistry B, 2006
A general approach is proposed to analyze electrical conductivities in aqueous solutions of polyb... more A general approach is proposed to analyze electrical conductivities in aqueous solutions of polybasic organic acids. Experimental conductivities are examined in the context of dissociation and hydrolysis reactions by applying the Quint-Viallard conductivity equations and the Debye-Hückel equations for activity coefficients. The proposed numerical procedure is illustrated by the case of benzenehexacarboxylic (mellitic) acid and its neutral and acidic salts. From conductivity measurements of mellitic acid and its salts, performed in dilute aqueous solutions in the 278.15-308.15 K temperature range, the limiting conductances of mellitic anions, lambda(0)(1/jH(6-j)Mel(-j), T), j = 1, 2, 3, 4, 5, 6 are determined.
The Journal of Physical Chemistry B, 2007
Systematic determinations of electrical conductivities of sodium penicillin G, potassium penicill... more Systematic determinations of electrical conductivities of sodium penicillin G, potassium penicillin G, and potassium penicillin V in the 278.15-313.15 K temperature range are reported. These conductivities are examined by applying the Quint-Viallard conductivity equations and the Debye-Hückel equations for activity coefficients. Determined dissociation constants and the limiting conductances of penicillin anions are based on the assumption that in dilute aqueous solutions, penicillin salts behave as acidic salts of dibasic acids, which are the final products of degradation reactions in acidic media. conductivity equations, 39-42 taking into account the dissociation of likely products of penicillin hydrolysis and degradation reactions.
PLoS ONE, 2012
Background: ATP-dependent D-alanine:D-alanine ligase (Ddl) is a part of biochemical machinery inv... more Background: ATP-dependent D-alanine:D-alanine ligase (Ddl) is a part of biochemical machinery involved in peptidoglycan biosynthesis, as it catalyzes the formation of the terminal D-ala-D-ala dipeptide of the peptidoglycan precursor UDPMurNAcpentapeptide. Inhibition of Ddl prevents bacterial growth, which makes this enzyme an attractive and viable target in the urgent search of novel effective antimicrobial drugs. To address the problem of a relentless increase in resistance to known antimicrobial agents we focused our attention to discovery of novel ATP-competitive inhibitors of Ddl.
Monatshefte für Chemie - Chemical Monthly, 2011
Abstract The density, refractive index, and electrical permittivity of 1,4-dioxane solutions of ... more Abstract The density, refractive index, and electrical permittivity of 1,4-dioxane solutions of acesulfame [6-methyl-1,2,3-oxathiazine-4(3H)-one-2,2-dioxide] and saccharin [1,2-benzisothiazole-3(2H)-one-1,1-dioxide] were measured at 298.15 K. From the experimental data the limiting apparent specific volume, refraction, and polarization of acesulfame and saccharin were calculated. The electrical dipole moments of acesulfame and saccharin were estimated according to the Debye, Onsager, and Kirkwood approaches. The association via dipole–dipole
Langmuir, 2012
A systematic investigation of the micellization process of a biocompatible zwitterionic surfactan... more A systematic investigation of the micellization process of a biocompatible zwitterionic surfactant 3-[(3-cholamidopropyl)dimethylammonium]-1-propanesulfonate (CHAPS) has been carried out by isothermal titration calorimetry (ITC) at temperatures between 278.15 K and 328.15 K in water, aqueous NaCl (0.1, 0.5, and 1 M), and buffer solutions (pH = 3.0, 6.8, and 7.8). The effect of different cations and anions on the micellization of CHAPS surfactant has been also examined in LiCl, CsCl, NaBr, and NaI solutions at 308.15 K. It turned out that the critical micelle concentration, cmc, is only slightly shifted toward lower values in salt solutions, whereas in buffer media it remains similar to its value in water. From the results obtained, it could be assumed that CHAPS behaves as a weakly charged cationic surfactant in salt solutions and as a nonionic surfactant in water and buffer medium. Conventional surfactants alike, CHAPS micellization is endothermic at low and exothermic at high temperatures, but the estimated enthalpy of micellization, ΔH M 0 , is considerably lower in comparison with that obtained for ionic surfactants in water and NaCl solutions. The standard Gibbs free energy, ΔG M 0 , and entropy, ΔS M 0 , of micellization were estimated by fitting the model equation based on the mass action model to the experimental data. The aggregation numbers of CHAPS surfactant around cmc, obtained by the fitting procedure also, are considerably low (n agg ≈ 5 ± 1). Furthermore, some predictions about the hydration of the micelle interior based on the correlation between heat capacity change, Δc p,M 0 , and changes in solvent-accessible surface upon micelle formation were made. CHAPS molecules are believed to stay in contact with water upon aggregation, which is somehow similar to the micellization process of short alkyl chain cationic surfactants.
Langmuir, 2013
Specific effects of the sodium salts of m- and p-hydroxybenzoates (m-HB and p-HB) on the aggregat... more Specific effects of the sodium salts of m- and p-hydroxybenzoates (m-HB and p-HB) on the aggregation process of dodecyltrimethylammonium chloride have been investigated by isothermal titration calorimetry, electrical conductivity, and (1)H NMR and compared with already reported data for the sodium salt of o-hydroxybenzoate (o-HB). For p-HB, it has been found that the aggregate is only formed by spherical micelles at all p-HB concentrations. On the other side, the situation is more complex for o-HB, where two distinct states of aggregation can be involved, depending on the concentration of o-HB. At high salt concentration, rodlike micelles are formed, whereas at lower concentration spherical aggregates are predominant. The transition from the cylinder to the sphere increases the mobility of the surfactant because the core of the rodlike micelles is more closely packed due to the expulsion of water from the interior of the aggregate. m-HB exhibits an intermediate behavior between these two extreme situations. The effect of the position of hydrophilic substituents on the aromatic ring on the insertion of the hydroxybenzoate anion in the micellar aggregate has been discussed.
Journal of Solution Chemistry, 2008
The electrical conductivity of aqueous cyclamic acid was studied in the concentration range (0.00... more The electrical conductivity of aqueous cyclamic acid was studied in the concentration range (0.0004 < c/mol·dm −3 < 0.14) at temperatures ranging from 278.15 to 303.15 K. Conductivities were measured by a precise method and examined by applying extended conductivity equations taking into account dimerization and incomplete electrolyte dissociation. Limiting molar electrolyte and ionic conductivities, dissociation and dimerization constants, and thermodynamic functions associated with dissociation and dimerization of cyclamic acid are discussed.
Journal of Solution Chemistry, 2008
The viscosities of aqueous solutions of lithium, sodium, potassium, rubidium and caesium cyclohex... more The viscosities of aqueous solutions of lithium, sodium, potassium, rubidium and caesium cyclohexylsulfamates were measured at 293.15, 298.15, 303.15, 313.15 and 323.15 K. The relative viscosity data were analyzed and interpreted in terms of the Kaminsky equation, η r = 1 + Ac 1/2 + Bc + Dc 2 . The viscosity A-coefficient was calculated from the Falkenhagen-Dole theory. The viscosity B-coefficients are positive and relatively large. Their temperature coefficient ∂B/∂T is negative or near zero for lithium and sodium salts whereas for potassium, rubidium and caesium salts it is positive. The viscosity D-coefficient is positive. This was explained by the size of the ions, structural solute-solute interactions, hydrodynamic effect, and by higher terms of the long-range Debye-Hückel type of forces. From the viscosity B-coefficients the thermodynamic functions of activation of viscous flow were calculated. The limiting partial molar Gibbs energy of activation of viscous flow of the solute was divided into contributions due to solvent molecules and the solute in the transition state. The activation energy of the solvent molecules was calculated using the limiting Gibbs energy of activation for the conductance of the solute ions. The activation energy of the solvent molecules was then discussed in terms of the nature of the alkali-metal ions and their influence on the structure of water. The limiting activation entropy and enthalpy of the solute for activation of viscous flow were interpreted by ion-solvent bond formation or breaking in the transition state of the solvent. The hydration numbers of the investigated electrolytes were calculated from the specific viscosity of the solutions.
Journal of Solution Chemistry, 2005
The electric conductivities of aqueous solutions of the lithium, sodium, potassium and ammonium s... more The electric conductivities of aqueous solutions of the lithium, sodium, potassium and ammonium salts of cyclohexylsulfamic acid were measured from 5 to 35 • C (in steps of 5 • C) in the concentration range 3 × 10 −4 < c/mol-dm −3 < 0.01. Data analysis based on a chemical model of electrolyte solutions yielded the limiting molar conductance ∞ and the association constant K A . Using the known values of the limiting conductances of lithium, sodium and potassium ions, the limiting conductances of the cyclohexylsulfamic ion were evaluated. Total dissociation of the investigated salts in water and negligible hydration of the cyclohexylsulfamate anion are evident.