Anda di halaman 1dari 6

P R E P R I N T ICPWS XV Berlin, September 811, 2008

UV-visible Spectroscopic Study on Nitrophenols Ionization Reactions to 225 oC


Jana Ehlerovaa, Liliana Trevanib, Josef Sedlbauera, Karine Ballerat-Busserollesc and Peter R. Tremaineb Department of Chemistry, Technical University of Liberec, 46117 Liberec, Czech Republic b Department of Chemistry, University of Guelph, Guelph, Ontario, Canada N1G 2W1 c Lab. de Thermodyn. des Solutions et des Polymres, Universit Blaise Pascal, 63177 Aubire, France Email: jana.ehlerova@tul.cz The UV-visible spectra of aqueous o-, m-, and p-nitrophenol were measured as a function of pH at temperatures from 50C to 225C at a pressure of 7 MPa. These were used to determine equilibrium constants for the acid ionization reaction of each isomer. The new experimental results, along with data from the literature, were used to develop a thermodynamic model to describe the dependence of ionization properties on temperature and pressure. The model yields predictions of the ionization constants for o-, m-, and pnitrophenol, log Ka, to at least 250C and 20 MPa with an estimated uncertainty in log Ka of less than 0.06. Thermal decomposition studies, carried out by using the system as a stopped-flow reactor, showed that the three isomers are significantly more stable in HCl acid solutions. At temperatures above 225 oC and alkaline solutions, the residence time of the solution in the cell and the pre-heater is so long that significant changes in the spectra of the nitrophenolate species are observed in stopped-flow experiments as an indication of thermal decomposition reactions.
a

Introduction Nitrophenols are used as raw materials in the chemical, pharmaceutical, dyes and herbicides industries, and are also produced in gas-phase oxidation of simple aromatic hydrocarbons [1]. Their toxicity is high even at trace levels, particularly for green plants and for aquatic organisms in surface waters. Thermodynamic properties are needed for the design of effective remediation technologies, for instance, steam oxidation. Nitrophenols are well-known colorimetric pH indicators, due to the differences in the UV-visible spectra of the protonated and unprotonated species (190 - 500 nm). However, only the ionization constants for p-nitrophenol have been determined at temperatures above 75 C by colorimetric methods [2]. The main objective of this work was to supplement the existing experimental information on aqueous nitrophenols (see for instance references in [3]) by determining their acid-base ionization constants over a wide range of temperature. A second objective was to develop a thermodynamic model able to describe these reactions in a form that is suitable for developing predictive models, such as functional-group contribution models, for describing the properties of structurally similar solutes in water.

Experimental Methods Ionization constants were measured by UVvisible spectroscopy in a high temperature-high pressure platinum flow cell with sapphire windows. A schematic diagram of the cell and injection system is shown in Figure 1. Details of the construction are described elsewhere [4]. Ammonia/ammonium buffer solutions were used to prepare nitrophenol solutions at fixed and known pH, with different ratios of protonated to unprotonated species. To determine the spectra of the fully protonated and fully unprotonated indicator, nitrophenol was dissolved in solutions of dilute HCl(aq), (the acid extreme) and NH3(aq) or

Figure 1: Schematic of high temperature-high pressure cell and injection system.

NaOH(aq) (the base extreme). For o- and pnitrophenol at 225 oC and NaOH(aq) (the base extreme) for m-nitrophenol and p-nitrophenol at temperatures under 200 oC. Thermal Stability Experiments Stopped-flow experiments were conducted at 25 and 225C in order to study the thermal stability of nitrophenols in the aqueous ammonia/ammonium buffer solutions and in the solutions of HCl(aq), NaOH(aq) and NH3(aq) used to define the two extremes. In these experiments the flow was stopped for 10 minutes and the absorbance spectra between 250 and 500 nm were measured every 30 seconds. Significant time-dependant changes in absorbance were observed only for the o- and pnitrophenol solutions in aqueous NaOH at 225 oC, but not for m-nitrophenol. When NH3 was used instead of NaOH, the changes were small and compatible with the time scale of the experiments (collection times). Typical results are shown in Figures 2 and 3. No attempt was made to determine the decomposition mechanism in alkaline media or to identify the decomposition products. Although the uncertainty associated with experiments at 225 oC is somewhat larger than that at lower temperatures, the experimental results at this temperature were included in the model presented below. Spectroscopic analysis UV-visible spectra for the solutions of o-, mand p-nitrophenol at constant pressure, 7 MPa, and temperatures from 100 C to 225 C were collected at wavelength intervals of 0.25 nm from 200 nm to 800 nm and a scan rate of 1200 nm.min-1. Representative spectra for all three isomers at 175 o C are shown in Figure 4. The temperature range was determined by the useful pH range of the ammonia/ ammonium buffer solutions which defined the low temperature limit, and the thermal stability of the indicator. At temperatures over 225C, the results may include systematic errors because of thermal decomposition of the indicator. The deconvolution of the nitrophenol spectra in the buffer solutions was carried out by linear combination of the spectra of the acid and the base forms of the nitrophenol species (Beers law). From the deconvolution of the spectra obtained for each buffer system, the concentration ratio of ionized and un-ionized forms of the nitrophenol 2

Figure 2: Stopped-flow experiment for o-nitrophenol in alkaline media at 225 C and 7 MPa : (a) in NaOH(aq) and (b) in aqueous NH3(aq). Time between spectra, 30 seconds.

Figure 3: Stopped-flow experiments for m-nitrophenol in NaOH(aq) base extreme and pnitrophenol in NH3 (aq) base extreme at 225 C and 7 MPa. Time between spectra, 30 seconds.

(the indicator ratio) could be determined at known pH. These ratios could then be used to calculate the ionization constant. The resulting values of the fitted ionization constants for all the buffer solutions were averaged to determine the experimental pKa value at each temperature. The number of buffer solutions at each temperature ranged from 2 to 6. Results Average pKa values for the three nitrophenol isomers are tabulated elsewhere [3], along with their estimated uncertainties. The uncertainty estimates are composed of the standard deviation of the calculated pKas at a given temperature plus the contribution of possible systematic errors in solution preparation. The two uncertainty contributions are usually comparable and do not exceed 0.03 pKa units each, leading to total uncertainties of up to 0.06 pKa units.

and calculated values, respectively, and X is the estimated experimental uncertainty.

pK aexp pK acalc F = + pK a i =1 i
O

H aexp H acalc + + H a j =1 j
P calc exp calc R C exp p,a C p,a + Va Va + l =1 C p,a Va k =1 l k Q 2 2

Thermodynamic model Our experimental values of pKa can be combined with other thermodynamic data for nitrophenol ionization taken from the literature. These include standard partial molar enthalpies H, volumes V, and heat capacities Cpwhich are known near 25 C. Together with our results, there are enough experimental results to develop a model describing the changes of Ka for each nitrophenol with temperature and pressure. Several equations of state for the standard thermodynamic properties of hydrothermal solutions could be applied for this purpose [e.g. 5, 6, 7]. The model proposed originally by Marshall and Franck [6] as a correlation tool for the ionization constant of water was chosen, because of its simplicity. For Ka the equation takes the form,

ln K a = pK a . ln 10 = =a+ b c d f g + 2 + 3 + e + + 2 . ln w T T T T T

where w is the density of water, T is thermodynamic temperature, and a g are fitting parameters. Other ionization properties are obtained by appropriate derivations of this equation. The ionization data for each nitrophenol were subjected to a simultaneous weighted fit with the objective function where exp and calc stand for experimental 3

Figure 4 : Normalized spectra at t = 175 C, p = 7 MPa for aqueous o-, m-, and p-nitrophenol. The arrows indicate the direction of increasing pH. The resulting fitted parameters of the MarshallFranck model for the three nitrophenols are reported in [3]. Figure 5 provides a graphical comparison of experimental Ka with predictions of Marshall and Franck model along the saturation line for water. The fit of the model to the experimental data was well within the estimated uncertainties of all the data used in the regression. The only experimental values for comparison with our results are the ionization constants of p-nitrophenol reported by Shin et al. [2] at temperatures up to 200C. These were obtained by UV-visible spectroscopy using methods similar to ours in batch cells, rather than flow cell. The results agree to within the combined estimated uncertainties, although Shin et al.s results for pKa for pnitrophenol are systematically more positive with increasing temperature below 175 C. We regard their measurement at 200 C as suspect, because of the possibility of thermal decomposition, which should be more pronounced in a batch cell due to the much longer exposure times. Possible reasons for the systematic difference between the two studies include the choice of activity coefficients, and approximations used in data treatment. Unfortunately, it is not clear from the Shin et al. paper, how the solutions were prepared and what was the target ionic strength, so we could not reconstruct their treatment precisely. Solvation Effects The effects of temperature and pressure on solvation are reflected in the standard state thermodynamic properties derived from the model [6], see Fig.6. In our judgment, the database from which the model parameters were fitted is too limited to permit realistic estimates of the uncertainties associated with these calculated derivative functions by error propagation analysis. Instead, approximate uncertainty limits were estimated from the relative error of model predictions from experimental pKa data, and the magnification of uncertainty upon derivation, to determine the uncertainties in other properties [3]. This leads to estimated uncertainties in Ha of 0.4 - 0.8 kJmol-1; Sa, 3 - 6 JK-1mol-1; Cp,a , 10 - 40 JK-1mol-1; and Va, 0.4 - 4 cm3mol-1 where the lower limits are for 25C, the upper limits for 200C. The large uncertainty in Va at elevated 4

temperatures arises because with the exception of the volumetric data at 25C, the calculated values are entirely

Figure 5: Ionization constants of nitrophenols corrected to psat, plotted as pKa vs. T. Fitted Marshall and Franck model (full line). pKa data from this work corrected from p = 7 MPa to psat (>); o-nitrophenol from Judson and Kilpatrick [8] ( ); m-nitrophenol from Robinson and Peiperl [9] (); p-nitrophenol from Judson and Kilpatrick [8] ( ); p-nitrophenol from Shin et al. [2] () temperature dependence, rather than the pressure dependent on parameters determined by fitting the dependence of pKa. There are significant differences in the properties of the three isomers at room temperature. These have been attributed to the electronwithdrawing effects of the nitro group, which stabilize the nitrophenolate anion, and to solvation effects that manifest themselves in the entropy, heat capacity and volume of ionization. The electronwithdrawing effects are larger for the o- and pisomers than for m-nitrophenol (See, for example,

Dewick [11]). Fernandez and Hepler [12] have noted that m-nitrophenol is a weaker acid than o- or p-nitrophenol at room temperature because its entropy of ionization is significantly more negative, and not because of a more endothermic enthalpy of ionization. Apparent molar volume and heat capacity measurements on p- and m-nitrophenol and their sodium salts [13, 14] suggest that this effect arises primarily because of differences in solvation for the uncharged species, with V(pNphOH) < V(m-NphOH), and Cp(p-NphOH) < Cp(m-NphOH). These observations are all consistent with solvation effects in which water is more tightly bound by the o- and p- isomers than by the m- isomer at 25C. Although the relative positions of the Gibbs energies of ionization of the isomers do not change with increasing temperature, there are changes in the contributions of the standard state enthalpy and entropy. While both Ha and Sa become increasingly negative with increasing temperature, the entropic term is the major contribution to the Gibbs energy of ionization above about 200C, i.e. the absolute values of TSa >>Ha. This effect has been observed for a great many acids and bases [15, 16]. It arises from long-range polarization of water by the charged anion and hydrated proton, an effect which becomes much more important at elevated temperature as the dielectric constant of water decreases. While the relative order of Ha and Sa for the three isomers at temperatures above 200C is the same as that at room temperature, Sa(m-NphOH) < Sa(p-NphOH) Sa(o-NphOH), the relative position of the orthoisomer at intermediate temperature changes significantly, so that Sa(m-NphOH) < Sa(oNphOH) < Sa(p-NphOH). This effect, which lies just outside the combined estimates of estimated uncertainty, may reflect the competing effects of localized hydrogen bonding about the uncharged species and solvent polarization effects. Temperatures in the range t 100C to t 250C are typically considered a transition region, in which water behaves like a normal hydrogenbonded solvent, so that both effects are often of similar magnitude. Functions of the standard partial molar heat capacity and volume functions, Cp,a and Va, show tendencies that are associated with solvent polarization effects. There is the transition with increasing temperature from the distinctly different experimental values for each isomer at ambient temperatures, through a maximum, towards the large negative values associated with solvent 5

polarization in the near-critical region [15,16]. Because Cp,a and Va are derivative functions,

Figure 6: Standard partial molar heat capacity, Cp,a, and standard Gibbs energy of ionization, Ga, of nitrophenols as a function of temperature at psat. Full line: o-nitrophenol; long-dashed line: mnitrophenol; short-dashed line: p-nitrophenol there is considerable uncertainty in the hightemperature values, so that the results for the three isomers above ~ 150C agree to within the rather large combined error limits. Measurements are under way to determine the standard partial molar heat capacity and volume of each isomer and its conjugate base, in order to confirm whether significant differences do exist at elevated temperatures. Acknowledgement This work was supported by the International Association for the Properties of Water and Steam (IAPWS Young Scientist Fellowship for J.E.); the Natural Sciences and Engineering Research Council of Canada (NSERC); and by the Research Centre of Advanced Remedial Technologies and Processes at the University of Liberec. Literature

[1] J. Tremp, P. Mattrel, S. Fingler and W. Giger: Phenol and nitrophenol as tropospheric pollutants: emissions from automobile exhaust and phase transfer in the atmosphere. Water Air Soil Pollut. 68, 113123 (1993) [2] T.-W. Shin, K. Kim and I.-J. Lee: Spectrophotometric determination of the acid dissociation constants for cacodylic acid and pnitrophenol at elevated temperatures. J. Sol. Chem., 26, 379-390 (1997) [3] J. Ehlerova, L. Trevani, J. Sedlbauer and P.R. Tremaine: Spectrophotometric determination of the ionization constants of aqueous nitrophenols at temperatures up to 225 C. J. Solution. Chem., 37, 857-874 (2008) [4] E. Bulemela, L. Trevani and P.R. Tremaine: Ionization constants of aqueous glycolic acid at temperatures up to 250 C using hydrothermal pH indicators and UV-visible spectroscopy. J. Sol. Chem., 34, 769-787 (2005) [5] J.C. Tanger and H.C. Helgeson: Calculation of the thermodynamic and transport properties of aqueous species at high pressures and temperatures: Revised equations of state for the standard partial molal properties of ions and electrolytes. Am. J. Sci. 288, 1998 (1988) [6] W.L. Marshall and E.U. Franck: Ion product of water substance, 01000 C, 110,000 bars. New international formulation and its background. J. Phys. Chem. Ref. Data, 10, 295304 (1981) [7] J. Sedlbauer, J.P. OConnell and R.H. Wood: A new equation of state for correlation and prediction of standard molal thermodynamic properties of aqueous species at high temperatures and pressures. Chem. Geol. 163, 4363 (2000) [8] C. M. Judson and M. Kilpatrick: The effects of substituents on the dissociation constants of substituted phenols. I. Experimental measurements in aqueous solutions. J. Am. Chem. Soc. 71, 31103115 (1949) [9] R. A. Robinson and A. Peiper: The ionization constant of m-nitrophenol from 5 to 50 C. J. Phys. Chem. 67, 28602861 (1963) [10]K.S. Pitzer: In: Pitzer, K.S. (ed.) Activity Coefficients in Electrolyte Solutions, 2nd edn. CRC Press, Boca Raton (1991) [11]P.M. Dewick: Essentials of Organic Chemistry. Wiley, New York (2006). Chap. 4 [12]L.P. Fernandez and L.G. Hepler: Heats and entropies of ionization of phenol and some substituted phenols. J. Am. Chem. Soc. 81, 17831786 (1959) 6

[13]H.P. Hopkins, W.C. Duer and F.J. Millero: Heat capacity changes for the ionization of aqueous phenols at 25 C. J. Solution Chem. 5, 263268 (1976) [14]C.L. Liotta, A. Abidaud and H.P. Hopkins: Thermodynamics of acid-base equilibriums. IV. Influence of solvation factors on acidity. Volumes of ionization of the m- and p-isomers of nitrophenol and formylphenol in water at 25 o C. J. Am. Chem. Soc. 94, 8624 (1972) [15]R.E. Mesmer, W.L. Marshall, D.A. Palmer, J.M. Simonson and H.F. Holmes: Thermodynamics of aqueous association and ionization reactions at high temperatures and pressures. J. Solution Chem. 17, 699718 (1988) [16]P.R. Tremaine, P. Bnzeth, C. Xiao and K. Zhang, in: D.A. Palmer, R. Fernandez-Prini, A.H. Harvey (eds.) Aqueous Systems at Elevated Temperatures and Pressures: Physical Chemistry in Water, Steam and Aqueous Solutions, Chapt. 13. Elsevier, Amsterdam (2004)

Anda mungkin juga menyukai