Sei sulla pagina 1di 7

Environ. Sci. Technol.

1996, 30, 1198-1204

Modeling of Ion-Exchange Reactions on Metal Oxides with the Frumkin Isotherm. 1. Acid-Base and Charge Characteristics of MnO2, TiO2, Fe3O4, and Al2O3 Surfaces and Adsorption Affinity of Alkali Metal Ions
H I R O K I T A M U R A , * , NORIAKI KATAYAMA, AND RYUSABURO FURUICHI
Faculty of Engineering, Hokkaido University, Sapporo, 060 Japan, and Asahikawa National College of Technology, Asahikawa, 071 Japan

Introduction
Metal oxides and hydroxides are widespread in soils and sediments as small particles or colloids, and they exhibit large ion adsorption capacities due to their large surface areas. The ion adsorption ability of metal oxides arises from the acid-base surface hydroxyl groups formed by dissociative chemisorption of water molecules (1, 2). The protonation and deprotonation of surface hydroxyl groups produce electric charges on the oxide, and the charged surface adsorbs ions from the solution to maintain electric neutrality in both the solid and solution phases, that is, an ion exchange. The ion exchange, acid-base, and charge characteristics of oxides are important to environmental science and technology. In water treatment, metal oxides and hydroxides are used to remove hazardous ions from waters by adsorption, and if oxide particulate contaminants in waters are removed, they can be coagulated and settled by controlling the surface charge through ion adsorption. Radioactive waste from nuclear power plants is planned for disposal deep in the ground, and the possibility of underground migration of hazardous ions to the biosphere is a serious concern where adsorption by oxides as soil components is considered to play a determining role (35). Modeling of ion exchange on oxides is necessary to describe and predict the behavior of ions and oxide particles in environments. Also, a model would make it possible to design and control the processes of ion/oxide particle systems (waste disposal, water treatment, chemical analysis, and others) in more theoretical ways for better efficiencies. As a model, the mass action law is insufficient to reproduce the whole range of ion-exchange behavior. This is presumably because, unlike homogeneous reactions, the reaction products of ion exchange are localized to the solid/ solution interphase, and the lateral interaction between interphase species leads to suppression of further adsorption. Recently, surface complexation models have been developed with modified mass action law expressions, where the electric surface potential term is introduced and the lateral interaction can be described as an electrostatic repulsion (6-12). These models use the surface potential estimated from the structures and properties of the electrical double layer and appear to apply to ideal oxides with smooth surfaces, few pores, and well-defined compositions and structures to meet the double-layer assumptions. However, such ideal oxides are rarely seen in natural environments. Even with ideal oxides, metal oxide particles are difficult to make into electrodes, the surface potential cannot be measured or controlled directly, and it is difficult to substantiate the proposed electrical double-layer models (13-17). Furthermore, the models exclude nonelectric lateral interactions between interphase species, but it is very likely that geometric interactions, such as steric hindrance due to the size of the adsorbed ions that replace
* Corresponding author telephone: +81-11-706-6741; fax: +8111-706-7882. Hokkaido University. Asahikawa National College of Technology.

Metal oxides abundant in natural environments affect the concentrations of ions in waters by adsorption. The ion adsorption ability of oxides arises from the acid-base nature of surface hydroxyl groups formed by dissociative chemisorption of water molecules. The protonation and deprotonation reactions of hydroxyl groups produce electric charges, resulting in ion adsorption to maintain electric neutrality (ion exchange). The amount of surface charge in alkali metal nitrate solutions was measured as a function of pH by titration, and the ion-exchange reactions accompanying the charge formation were modeled by using the Frumkin isotherm, which assumes suppression of the reaction due to lateral interactions between the interphase species. This model embodies not only electrical but also chemical, geometrical, and/or other lateral interactions and can be applied to real, not well-defined oxide/solution systems in natural environments. From the model parameters, it was found that the intensity of cation exchange (deprotonation) increases in the order: Al2O3 < Fe3O4 < TiO2 < MnO2, and the intensity of anion exchange (protonation) decreases in the same order. The electronegativity of the lattice metal ions of these oxides was estimated and found to increase in the order above. It is suggested that with electronegativity of the lattice metal ions the electron density of adjacent lattice oxide ions and, hence the acid-base nature of hydroxyl sites, changes. Also, the adsorption affinity of alkali metal ions was evaluated and discussed within the model parameters.

1198

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 30, NO. 4, 1996

0013-936X/96/0930-1198$12.00/0

1996 American Chemical Society

from the charge separation (polarization) in the solid/ solution interphase. Material Balances. The acid and base surface hydroxyl group sites considered as independent species are equal in number (32), because they are formed from neutral water molecules. The material balance equations for surface sites are

[-OH(a)] + [-O-M+] ) Ns/2 [-OH(b)] + [-OH2+X-] ) Ns/2

(3) (4)

FIGURE 1. Acid and base hydroxyl sites on a metal oxide, and ion-exchange reactions at the oxide/solution interphase: (b) metal ions; (O) oxide ions; (a) acid hydroxyl sites; (b) base hydroxyl sites.

protons or hydroxide ions at surface sites, suppress ionexchange reactions (18). The authors have tried to model ion-exchange reactions on metal oxides and on a weak acid cation-exchange resin in different ways (19-29). In our model, the suppression of ion exchange due to lateral interactions is expressed by an increase in the Gibbs free energy change like the Frumkin isotherm, and this embodies all causes of suppression: electrical, chemical, geometrical, and/or others. As a consequence, the model can be applied flexibly to real oxides in soils and sediments. In this investigation, the model was applied to the ionexchange reactions on metal oxides with simple monovalent ions (alkali metal nitrates), where the action of ions is weak and the intrinsic properties of oxide surfaces are clearly observed. The extent of ion adsorption and the resulting surface charge were reproduced, and the different acidbase and ion-adsorption properties of oxides and affinities of alkali metal ions for an oxide were assessed and discussed within the model parameters.

where Ns is the surface density of all the hydroxyl sites (mol m-2) and half is the cation- and anion-exchange capacities. There is an alternative material balance equation assuming that the surface hydroxyl groups are amphoteric ([-OH] + [-O-M+] + [-OH2+X-] ) Ns). The modeling in this investigation can be equally well made with this equation, but the values of the model parameters become different from those with eqs 3 and 4. Equilibrium Conditions. The surface ion pairs are polarized, and their local charges may affect adjacent interphase species electrostatically. There will be a repulsion between charged sites and between adsorbed ions with the same sign. The H+ or OH- ions making up undissociated hydroxyl group sites will be attracted to the solid by the charges of neighboring dissociated hydroxyl sites of the same type. Further, the larger sizes of M+ and X- ions which replace H+ and OH- ions at hydroxyl sites may cause steric hindrances between adsorbed ions. All such lateral interactions, electrical, geometrical, chemical, and/or others suppress ion exchange and will become increasingly stronger as the surface coverage of exchange sites increases. Then, the Gibbs free energy changes in ion-exchange reactions 1 and 2 may be described by

Ga ) Ga + RT ln Qa + maa Gb ) Gb + RT ln Qb + mbb

(5) (6)

The Model
Stoichiometry. The cation- and anion-exchange reactions at acid and base hydroxyl group sites, -OH(a) and -OH(b), in (1:1) electrolyte, MX, solution can be expressed as

Here, Ga and Gb are the standard Gibbs free energy changes (energy mol-1); Qa and Qb are the concentration ratios:

-OH(a) + M+ a -O-M+ + H+ -OH(b) + H2O + X- a -OH2+X- + OH-

(1) (2)

Qa ) [-O-M+][H+]/([-OH(a)][M+]) Qb ) [-OH2+X-][OH-]/([-OH(b)][X-])

(7) (8)

Figure 1 illustrates the oxide/solution interphase where both reactions 1 and 2 are taking place. Here, the acid and base hydroxyl groups are considered as independent species located in different layers, as suggested by Schindler (1) and Stumm (2). The anions and cations are adsorbed electrostatically to the charged sites (nonspecific adsorption). These pairs of adsorbed ions and charged sites are termed surface ion pairs in the following. This molecular model is based on the following facts: (a) the stoichiometric ratio between released H+ and adsorbed M+ or released OH- and adsorbed X- is 1:1 (30); (b) by electrophoresis of the oxide particles, the surface with X- adsorbed is effectively positive and that with M+ is effectively negative (31); and (c) electric neutrality must be maintained in both the solid and solution phases. The surface charge is balanced by the counterions, but electrokinetic phenomena may occur in an electric field due to effective surface charges resulting

a and b are the coverages of the acid and base hydroxyl sites:

a ) [-O-M+]/(Ns/2) b ) [-OH2+X-]/(Ns/2)

(9) (10)

and ma and mb are the proportionality constants (energy mol-1). The linear terms with respect to the coverages express the effect of the above-mentioned lateral interactions; this assumption is identical to that for the Frumkin isotherm (33). The mutual interaction between the anionand cation-exchange reactions is not considered by eqs 5 and 6, and the reason for this will be discussed in the calculation of surface site densities section below.

VOL. 30, NO. 4, 1996 / ENVIRONMENTAL SCIENCE & TECHNOLOGY

1199

The equilibrium conditions of reactions 1 and 2 are derived from eqs 5 and 6 at Ga ) 0 and Gb ) 0:

exp(-Ga/RT) ) Ka ) Ka exp(Aaa) exp(-Gb/RT) ) Kb ) Kb exp(Abb)

(11) (12)

where Ka and Kb are the intrinsic equilibrium constants, Ka and Kb are the concentration ratios Qa and Qb at equilibrium, and Aa and Ab are the lateral interaction constants, being Aa ) ma/RT and Ab ) mb/RT. The equilibrium constants of ion-exchange reactions Ka and Kb can be regarded as the protonation/deprotonation constants of surface hydroxyl groups for fixed anion and cation concentrations. Ka and Kb correspond to the mass action law relations, but decrease with increasing because the exponential terms increase, while the intrinsic equilibrium constants must be unchanged in eqs 11 and 12. Hence, the exponential terms express the suppression of ion exchange by already formed surface ion pairs and are related to the activity coefficients of interphase species as will be described in the final section. The values of Aa and Ab are measures of the suppression effect due to the lateral interactions between interphase species. The suppression or lateral interaction constants Aa and Ab are dimensionless here, but in the previous paper (21), they were defined with respect to the surface density of exchanged sites. The model equations 11 and 12 are mathematically identical to the constant capacitance model, which assumes a specific type of electrical double layer. However, the model developed here is based on the Frumkin isotherm, which does not specify any particular physical or molecular features of the solid/solution interphase. The model covers not only electrical but also geometrical, chemical, and/or other lateral interactions. As a result, the lateral interaction constants Aa and Ab are not simply double-layer parameters and may be termed thermodynamic parameters.

FIGURE 2. Schematic titration curves of MOH-MNO3 solutions (A) with and (B) without oxide.

groups (-OH) react with CH3MgBr to evolve equivalent amounts of methane gas (CH4) (34):

-OH + CH3MgBr f -OMgBr + CH4v

(13)

Experimental Section
Materials. The manganese dioxides (MnO2) were International Common Samples (IC1, IC12, and IC22) for batteries, supplied from the IC Sample Office, Cleveland, OH. These were prepared by electrolytic (IC1) or chemical (IC12, IC22) oxidation of manganese ions, and the crystal structure is -type. Magnetite (Fe3O4) was a commercial reagent from Kanto Chemical Co., Tokyo, Japan. Titania (TiO2, rutile) and alumina (-Al2O3) were the reference catalysts, JRC-TIO-5 and JRC-ALO-4, respectively, supplied from the Catalysis Society of Japan. The manganese dioxide samples were washed with dilute nitric acid solution and then with distilled water until the washing water showed the pH of distilled water. The other samples were washed thoroughly with distilled water. The washed oxide powders were dried at 110 C overnight and kept in a silica gel desiccator. The specific surface area of samples, SBET (m2 g-1), was determined by the BET method with nitrogen gas adsorption. Determination of Surface Hydroxyl Groups. The amount of surface hydroxyl groups of metal oxides was determined by Grignard reaction of surface hydroxyl groups. A weighed portion of the oxide samples was put in dry ether solution containing excess methyl magnesium bromide (CH3MgBr), a Grignard reagent. The surface hydroxyl

The amount of CH4 evolved was measured with a gas buret, and the amount of -OH was determined. The surface concentration (density) of -OH, Ns (mol m-2), was obtained by dividing the amount of -OH by the surface area of the oxide samples. Determination of Surface Charge. The surface charge established on metal oxides by the release of H+ and OHions, as a result of ion exchange, was measured as a function of pH by acid-base titration. Weighed portions of metal oxide powders (0.4-11 g) were suspended in 200-cm3 solutions of 0.1 mol dm-3 alkali metal nitrates (MNO3), the suspension pH was 5-6, nitrogen was bubbled, and the suspensions were allowed to stand for 2-3 h to remove carbon dioxide completely. Fixed amounts of alkali metal hydroxides (MOH) were added, and the suspensions were titrated with nitric acid solutions with known concentrations at 25 C (Figure 2, curve A). The points in the titration curve were determined after the pH became stable, and in the slowest case, more than 3 h was allowed to assure the attainment of equilibrium. The anion X- used in this investigation is all nitrate ion. As a reference, titrations were also carried out for MOH-MNO3 solutions without oxides (Figure 2, Curve B). The surplus or deficiency in the amount of nitric acid added to the suspension with reference to the blank, x (mol), is given by the difference in the amounts of OH- and H+ ions, n(OH-rel) and n(H+rel), released by reactions 1 and 2:

x ) n(OH-rel) - n(H+rel) ) ([-OH2+NO3-] - [-O-M+])S ) S (14)


Here, n(OH-rel) and n(H+rel) are replaced by the equivalent amounts of surface ion pairs; the bracket terms are the surface ion-pair densities (mol m-2) and S is the surface area of the oxides (m2). The surface ion-pair density

1200

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 30, NO. 4, 1996

TABLE 1

Specific Surface Area SBET and Hydroxyl Site Density Ns of Metal Oxide Samples
sample MnO2-IC1 MnO2-IC12 MnO2-IC22 TiO2 Fe3O4 Al2O3

SBET (m2 g-1)


43.0 80.0 45.1 2.60 4.32 155

Ns (mol m-2)
2.25 10-5 2.35 10-5 2.41 10-5 1.80 10-5 3.31 10-5 3.20 10-5

difference is expressed by (mol m-2):

) [-OH2+NO3-] - [-O-M+]

(15)
FIGURE 3. Surface charge density vs pH for metal oxides at ionic strength 0.1 M (NaNO3) and 25 C: MnO2-IC1 (O), MnO2-IC12 (0), MnO2-IC22 (b), TiO2 (4), Fe3O4 (2), Al2O3 (9); points, experimental; curves, calculated.

This changes from negative to positive depending on pH and is equivalent to the net surface charge density in mol m-2 (if this is multiplied by the Faraday constant, F, the surface charge density in coulomb m-2 is obtained). The values of were obtained as x/S from titration curves. At the intersection of curves A and B, is zero, and this pH is the point of zero charge (pzc). The concentrations of ions remaining in the solution after adsorption can be regarded as equal to the initial concentrations, since the fraction of adsorbed ions was less than 2% under these conditions. Determination of Model Parameters. By fitting eqs 3, 4, 11, 12, and 15 to the measured , the values of Ka, Kb, Aa and Ab were determined by the procedure described elsewhere (21).

Results and Discussion


Specific Surface Areas and Surface Hydroxyl Site Densities. Table 1 shows the specific surface area SBET and the surface hydroxyl site density Ns of the metal oxide samples. The specific surface area changes with the preparation conditions and differs greatly from sample to sample. However, the Ns values are similar and can be regarded as close to the value calculated from the size of the hydroxide ion. The radius of the hydroxide ion is 1.45 (35), and the maximum density of hydroxyl groups in one layer with the closest packing would be 1.95 10-5 mol m-2. As indicated in Figure 1, the acid and base surface hydroxyl groups are located in two layers, and the total hydroxyl site density becomes 3.90 10-5 mol m-2, comparable to the measured Ns values in Table 1. Similar Ns values have been obtained by other methods as described by Westall and Hohl (13): 2.07 10-5 mol m-2 for TiO2 by the tritium-exchange method and crystallographic calculation (Yates); 2.07 10-5 mol m-2 for -Al2O3 by the crystallography (Peri); 1.41 10-5 mol m-2 for -Al2O3 by the tritium-exchange method (Kummert). However, Westall and Hohl (13) also quoted a far smaller value of 0.265 10-5 mol m-2 obtained for -Al2O3 by Hohl and Stumm with saturation of surface deprotonation, and the availability of all the crystallographic surface hydroxyl sites to the reaction was questioned. For this technique, it should be remembered here that ion adsorption (surface protonation/deprotonation) becomes increasingly suppressed with the progress of the reaction due to lateral interactions and that surface saturation is difficult to attain. We have reported that the complete deprotonation of acidtype surface hydroxyl groups on Fe2O3 requires very high concentrations of free OH- (>5 mol dm-3) (29), and so the
FIGURE 4. Surface charge density vs pH for MnO2-IC12 with different alkali metal ions (M+) at ionic strength 0.1 M (MNO3) and 25 C: Li+ (O), Na+ (0), K+ (4), Cs+ (b).

Ns value obtained for Al2O3, which is easily dissolved in alkaline solution, would not be reliable. Relationship between Surface Charge Density and pH. Figure 3 shows the relationship between the surface charge density and pH for different oxide samples in NaNO3 solutions. Positively charged sites with anions adsorbed are predominant at lower pH ( is positive), and negatively charged sites with cations adsorbed are predominant at higher pH ( is negative) for all the oxides, as predicted from the pH dependencies of reactions 1 and 2. The -pH curves are different for different oxide samples; the pzc (pH at ) 0) varies, and even for the same oxide (MnO2), it changes depending on the preparation conditions. The -pH relations for different alkali metal ions with MnO2-IC12 were also obtained, and the values at pHs more than 1 pH higher than pzc are plotted in Figure 4 (the curve for Na+ is the same as that in Figure 3). It is seen that at high pH the cation adsorption density increases in the order: Cs+ e K+ < Na+ < Li+. Model Parameters. The optimum values of the parameters obtained by fitting the model equations to the -pH data for different metal oxides in 0.1 mol dm-3 NaNO3 solutions (Figure 3) are shown in Table 2. The curves in Figure 3 are the results of the best fit, reproducing the measured data (points) well. The application of the model

VOL. 30, NO. 4, 1996 / ENVIRONMENTAL SCIENCE & TECHNOLOGY

1201

TABLE 2

Optimum Values of Model Parameters and pzc for Metal Oxide Samples at Ionic Strength 0.1 M (NaNO3) and 25 C
sample MnO2-IC1 MnO2-IC12 MnO2-IC22 TiO2 Fe3O4 Al2O3 log Ka -3.98 -4.25 -5.11 -6.10 -6.66 -9.25 log Kb -9.95 -10.9 -9.89 -9.20 -8.04 -6.05 log Aa 1.35 1.47 1.57 1.39 1.65 1.31 log Ab 0.96 0.91 0.99 1.63 1.76 1.90 pzc 4.15 3.76 4.72 5.36 6.25 8.44

equations to other ionic strengths was attempted with the NaNO3/MnO2-IC22 system, and it was shown that this model applies reasonably well to ionic strengths ranging from 0.05 to 1.0 mol dm-3 (21). The samples used in this investigation are real oxides, like MnO2 used in battery electrodes, Al2O3 and TiO2 used as catalysts, and Fe3O4, a commercial reagent. These are also common in soils and sediments, and it would be possible to use the model parameters determined here to characterize the real oxides important in natural environments and industry. The Ka and Kb values for MnO2 show some scatter depending on the method of sample preparation. However, different acid-base properties of oxide samples of the same kind are not surprising as shown in the variety of pzc values of oxides (36, 37), and the possible causes of the scatter will be discussed in the effect of electronegativity section below. For the acid (cation-exchange) properties of metal (hydr)oxides, MnO2 is one of the strongest oxides (38-40), and the Ka value order here coincides with this: Al2O3 < Fe3O4 < TiO2 < MnO2. The base (anion-exchange) intensity indicated by Kb, however, decreases in this order. The opposite orders and the ordering in the acid-base intensities of the oxides are interesting and seem to point to some physically meaningful properties of the oxides as will be discussed in the effect of electronegativity section. The same series is seen in the Ab values, while the differences in the Aa values are smaller. The -pH data obtained for different alkali metal ions with MnO2-IC12 (Figure 4) were analyzed by a graphical method in place of computation to demonstrate how the data fit to

FIGURE 5. Plot of log Ka against a for MnO2-IC12 with different alkali metal ions (M+) at ionic strength 0.1 M (MNO3) and 25 C: Li+ (O), Na+ (0), K+ (4), Cs+ (b).
TABLE 3

Optimum Values of Model Parameters for Alkali Metal Ions (M+) with MnO2-IC12 at Ionic Strength 0.1 M (MNO3) and 25 C
ion Li+ Na+ K+ Cs+ log Ka -4.50 -4.25 -3.40 -3.20 log Aa 1.29 1.47 1.65 1.71

log Ka ) log Ka - aAa log e

(16)

This equation is obtained by taking the log of eq 11, e being the base of the natural log. In this pH region (above 1 pH higher than pzc), it is likely that the cation-exchange reactions are predominant and the anion-exchange reactions are negligible due to the opposite pH dependencies of the reactions. Then, the absolute values of (||) were regarded as equal to [-O-M+]; the Ka and a values were calculated with the measured as Ka ) ||[H+]/([Ns/2 ||][M+]) and a ) ||/(Ns/2). They are plotted in Figure 5. The good linear relations between log Ka and a demonstrate good fit to eq 11, where Ka and Aa are assumed constant. The Ka and Aa values obtained from the intercept and slope of the straight lines are listed in Table 3; the values for Na+ agree with those determined by the computations (Table 2). The Ka value increases in the order: Li+ < Na+ < K+ < Cs+; the Aa value changes in the same order. The changes of the model parameters with the kind of alkali metal ions will be discussed in the adsorption affinity section below.

For the -pH data obtained at pHs more than 1 pH lower than pzc, good fit to eq 12 has also been demonstrated (23). Calculation of Surface Site Densities. With the established model parameters, the surface densities of exchanged and unexchanged sites on Fe3O4 (as an example) were obtained as a function of pH at an ion concentration of 0.1 mol dm-3 (NaNO3) by solving the model equations (Figure 6). The negative site density increases and the positive site density decreases with pH, and they intersect at pzc. The undissociated acid and base hydroxyl sites are predominant over a wide pH range. This model calculation enables the prediction of the extent of ion adsorption and the surface charge for given conditions. The anion- and cation-exchange reactions take place simultaneously, producing oppositely charged sites. One may think it necessary to evaluate the interaction between the anion- and cation-exchange reactions (mutual interaction), where one exchange reaction enhances another due to attraction between the oppositely charged sites. However, in pH regions outside pzc, one exchange reaction dominates due to the opposite pH dependencies of the reactions, and the effect of mutual interaction on experimentally measured seems small as discussed next. The surface charge density is the difference in the positive and negative site densities, and either of the two will dominate. The effect of mutual interaction on the dominating site from the minor site is small, and the mutual interaction has little effect on , and so cannot be assessed by analyzing . Equations 11 and 12 are, therefore, sufficient and adequate to analyze and reproduce the observed . The calculated densities of minor charged sites in Figure 6 are due to the assumption that there is no mutual

1202

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 30, NO. 4, 1996

FIGURE 6. Calculation of site density vs pH for Fe3O4 at ionic strength 0.1 M (NaNO3) and 25 C.

FIGURE 7. Relationship between log Ka or log Kb and electronegativity Xi of lattice metal ions: Ka (O), Kb (b).

interaction. To assess the minor charged site densities, experimentally undeterminable assumptions (e.g., the assumption of no interaction here) must be made, as are invoked implicitly in the EDL models. Modeling of pzc. At pzc, the negative and positive site densities are equal (o), and the following equations are derived:

pzc ) {pKa - pKb + pKw + 2(Aa - Ab)o log e/Ns}/2 (17)


where Kw is the ion product of water (10-14 at 25 C). Also

Kwo2 exp{2(Aa + Ab)o/Ns} ) KaKbI2(Ns/2 - o)2 (18)


where I is the ionic strength of solution equal to the anion and cation concentrations under the experimental condition here. Equation 18 can be solved for o, and by substituting this o into eq 17, the pzc can be calculated. The calculated pzc values are shown in Table 2, and they agree well with those obtained graphically from Figure 3, as seen in the comparison of calculated with measured (model parameter section). The differences in pzc with the kind of oxide samples can be discussed with the model parameters. Effect of Electronegativity of Lattice Metal Ions on Surface Acid-Base Intensities and Lateral Interactions. Tanaka and Ozaki (41) reported the following equation for the electronegativity, Xi, of the lattice metal ions of oxides:

Xi ) (1 + 2Z)Xo

(19)

where Z is the charge of the metal ions and Xo is the electronegativity of corresponding elemental metals. For the metal oxides examined in this investigation, the Xi values were calculated, and log Ka and log Kb were plotted against Xi (Figure 7). The charge of iron ions in Fe3O4 was assumed to be +2.7 as an average. It is seen that log Ka increases while log Kb decreases with Xi in a symmetrical way, showing good correlations. From Figure 1, it is likely that, with increasing electronegativity of lattice metal ions, the

electron density of adjacent lattice oxide ions inductively decreases, then the acid hydroxyl site more easily loses its proton, and the base hydroxyl site gains a proton (loses a hydroxide ion) with some difficulty. The MnO2 samples prepared under different conditions showed differences in the values of Ka and Kb, suggesting that factors other than the electronegativity (like the crystal structure, the bond length in the oxide lattice, the pore structure, and others) also affect the acid-base properties of oxide surfaces. The effect of such factors on the acidbase nature of an oxide can be studied with this model by characterizing samples with systematic changes of the factors. As mentioned above, the suppression constants in this model are not restricted to the electrostatic interaction alone. However, the changes in the suppression constants Aa and Ab with the kind of oxide (Table 2) seem to be caused by an induction effect from the lattice metal ions on the electric density of the charged sites. This is because the adsorbed ions are common to all the oxides here, and the steric repulsions must be similar. The charge density of -OH2+ sites would increase with increasing electronegativity of adjacent lattice metal ions; this reduces the charge separation between the surface site and the adsorbed anion due to increasing attraction. The lateral repulsion (Ab) would then decrease in the direction seen in Table 2. The Aa values may also be explained in this manner. In other investigations, pzc has been considered to indicate the acid-base nature of metal oxides and has been examined as a function of the electronegativity (41) or the valence and the ionic radii (36, 42, 43) of lattice ions. However, pzc values are a composite of the parameters (Ka, Kb, Aa, and Ab) as described above, and the analysis of the respective parameters in this investigation seems to provide more detailed information of the acid-base nature of oxide surfaces. Adsorption Affinity of Alkali Metal Ions. With the results here, the adsorption affinity of the alkali metal ions evaluated from Ka increases with the atomic number of alkali metal ions, and the lateral repulsion evaluated from Aa also increases with the atomic number (Table 3). These

VOL. 30, NO. 4, 1996 / ENVIRONMENTAL SCIENCE & TECHNOLOGY

1203

indicate that the electric charge density of adsorbed alkali metal ions increases with the atomic number, when nonelectric lateral repulsion does not change with the kind of alkali metal ions. It is possible that the adsorbed ions are hydrated, and then the hydrated radius decreases in the atomic number order (44), while the charge density of hydrated ions increases in the same order. The selectivity series, Li+ < Na+ < K+ < Cs+, has been reported for manganese oxide in acid media (39), and this agrees with the Ka order. However, the observed order of adsorption densities at high pH (Figure 4) appears not to agree with the Ka order. Figure 5, a replot of the data in Figure 4, indicates that the log of concentration ratio (Ka) decreases with coverage (a) with different slopes (Aa), resulting in a reversal of the order of concentration ratios. It is likely that, with increasing coverage, the suppression (Aaa) increases, and ions with lower affinity (Ka) but lower repulsion (Aa) become adsorbed more easily than ions with higher affinity but higher repulsion. In acid media, the lateral interaction is small due to low coverage, and hence the selectivity series agreeing with Ka was obtained. Activity Coefficients of Interphase Species. The intrinsic equilibrium constants Ka and Kb can be defined with the activities of the species taking part in the reactions like

Ka ) Ka

y(-O-M+)y(H+) y(-OH(a))y(M+)

(20)

Kb ) Kb

y(-OH2+X-)y(OH-) y(-OH(b))y(X-)

(21)

where y represents the activity coefficient of the respective species. The ratios y(H+)/y(M+) and y(OH-)/y(X-) can be regarded as equal to 1 or constants close to 1 for a constant ionic strength. Combining these equations with eqs 11 and 12 gives

exp(Aaa) ) y(-O-M+)/y(-OH(a)) exp(Abb) ) y(-OH2+X-)/y(-OH(b))

(22) (23)

The exponential functions expressing the suppression of ion exchange are related to the activity coefficients of interphase species. The theory of the activity coefficients of interphase species has not been established. We have suggested the formulas of the activity coefficients y of the interphase species on an ion-exchange resin (27), and y(O-M+), y(-OH2+X-), y(-OH(a)), and y(-OH(b)) here can also be formulated in a similar way. Although it is difficult to establish the respective activity coefficients of interphase species numerically, the activity coefficient ratios (exponential terms) are determined in our modeling of interphase equilibria.

Acknowledgments
This work was supported by a grant-in-aid for Scientific Research from the Ministry of Education, Science and Culture, Japan (05680438).

(4) Combes, J.-M.; Chisholm-Brause, C. J.; Brown, G. E., Jr.; Parks, G. A.; Conradson, S. D.; Eller, P. G.; Trlay, I. R.; Hobart, D. E.; Meijer, A. Environ. Sci. Technol. 1992, 26, 376-382. (5) Bradbury, M. H.; Bayens, B. J. Colloid Interface Sci. 1993, 158, 364-371. (6) James, R. O.; Parks, G. A. In Surface and Colloid Science; Matijevic, E.; Ed.; Plenum: New York, 1982; Vol. 12, Chapter 2. (7) Westall, J. C.; Bousse, L.; Meindl, J. D.; Chan, D. Y. C.; Hayes, K. F.; Leckie, J. O.; Honeyman, B. D.; Leckie, J. O.; Yasunaga, T.; Ikeda, T. In Geochemical Processes at Mineral Surfaces; Davis, J. A., Hayes, K. F., Eds.; American Chemical Society: Washington, DC, 1986; ACS Symposium Series 323; Chapters 4-7, 9, and 12. (8) Westall, J. C.; Schindler, P. W.; Stumm, W. In Aquatic Surface Chemistry; Stumm, W., Ed.; John Wiley & Sons: New York, 1987; Chapters 1 and 4. (9) Dzombak, D. A.; Morel, F. M. M. Surface Complexation Modeling; John Wiley & Sons: New York, 1990; Chapter 2. (10) Davis, J. A.; Kent, D. B.; Sposito, G.; Schindler, P. W. In MineralWater Interface Geochemistry; Hochella, M. F., Jr., White, A. F., Eds.; Mineralogical Society of America: Washington, DC, 1990; Reviews in Mineralogy, Vol. 23; Chapters 5-7. (11) Stumm, W. Chemistry of the Solid-Water Interface; John Wiley & Sons: New York, 1992; Chapter 3. (12) Morel, F. M. M.; Hering, J. G. Principles and Applications of Aquatic Chemistry; John Wiley & Sons: New York, 1993; Chapter 8. (13) Westall, J.; Hohl, H. Adv. Colloid Interface Sci. 1980, 12, 265294. (14) Sposito, G. J. Colloid Interface Sci. 1983, 91, 329-340. (15) Johnson, R. E., Jr. J. Colloid Interface Sci. 1984, 100, 540-554. (16) Noh, J. S.; Schwarz, J. A. J. Colloid Interface Sci. 1990, 139, 139148. (17) Goldberg, S. J. Colloid Interface Sci. 1991, 145, 1-9. (18) Jin, X.; Talbot, J.; Wang, N.-H. L. AIChE J. 1994, 40, 1685-1696. (19) Tamura, H.; Matijevic, E.; Meites, L. J. Colloid Interface Sci. 1983, 92, 303-314. (20) Tamura, H.; Katayama, N.; Furuichi, R. Bunseki Kagaku 1988, 37, 395-399. (21) Tamura, H.; Oda, T.; Nagayama, M.; Furuichi, R. J. Electrochem. Soc. 1989, 136, 2782-2786. (22) Tamura, H.; Tatsumi, T.; Furuichi, R. Hyomen Gijyutsu 1989, 40, 1116-1120. (23) Tamura, H.; Furuichi, R. Bunseki Kagaku 1991, 40, 635-640. (24) Tamura, H.; Oda, T.; Furuichi, R. Anal. Chim. Acta 1991, 244, 275-280. (25) Katayama, N.; Tamura, H.; Furuichi, R. Denki Kagaku 1992, 60, 887-892. (26) Tamura, H.; Katayama, N.; Furuichi, R. Bunseki Kagaku 1993, 42, 719-724. (27) Tamura, H.; Kudo, M.; Furuichi, R. Anal. Chim. Acta 1993, 271, 305-310. (28) Katayama, N.; Tamura, S.; Tamura, H.; Furuichi, R. Denki Kagaku 1994, 62, 251-256. (29) Tamura, H.; Ohkita, K.; Katayama, N.; Furuichi, R. Bunseki Kagaku 1994, 43, 831-836. (30) Gohsh, B.; Chakravarty, S. N.; Kundu, M. L. J. Indian Chem. Soc. 1951, 28, 319. (31) Murray, J. W. Geochim. Cosmochim. Acta 1975, 39, 635. (32) Boem, H. P. Discuss. Faraday Soc. 1971, 52, 264. (33) Gileade, E.; Kirowa-Eisner, E.; Penciner, J. Interfacial Electrochemistry; Addison-Wesley: Reading, MA, 1975; pp 82-83. (34) Fripiat, J. J.; Uytterhoeven, J. J. Phys. Chem. 1962, 66, 800-805. (35) Kiriyama, R.; Kiriyama, H. Kozo Muki Kagaku; Kyoritsu Syuppan: Tokyo, 1964; p 283. (36) Parks, G. A. Chem. Rev. 1965, 65, 177-198. (37) Mushiake, K.; Masuko, N. Seisan Kenkyu 1977, 29, 2-10. (38) Amphlett, C. B. Inorganic Ion Exchangers; Elsevier: Amsterdam, 1964; p 86. (39) Fuller, M. J. Chromatogr. Rev. 1971, 14, 45-76. (40) Abe, M. Bunseki Kagaku 1974, 23, 1254-1285. (41) Tanaka, K.; Ozaki, A. J. Catal. 1967, 8, 1-7. (42) Carre, A.; Roger, F.; Varinot, C. J. Colloid Interface Sci. 1992, 154, 174-183. (43) Bleam, W. F. J. Colloid Interface Sci. 1993, 159, 312-318. (44) Cotton, F. L.; Wilkinson, G. W. Advanced Inorganic Chemistry; John Wiley & Sons: New York, 1966; p 422.

Literature Cited
(1) Schindler, P. W. In Aquatic Surface Chemistry; Stumm, W., Ed.; John Wiley & Sons: New York, 1987; p 87. (2) Stumm, W. Chemistry of the Solid-Water Interface; John Wiley & Sons: New York, 1992; p 13. (3) Girvin, D. C.; Ames, L. L.; Schwab, A. P.; McGarrah, J. E. J. Colloid Interface Sci. 1991, 141, 67-78.

Received for review June 16, 1995. Revised manuscript received November 3, 1995. Accepted November 7, 1995.X ES9504404
X

Abstract published in Advance ACS Abstracts, February 1, 1996.

1204

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 30, NO. 4, 1996

Potrebbero piacerti anche