You are on page 1of 10

Journal of Colloid and Interface Science 303 (2006) 214–223

www.elsevier.com/locate/jcis

Parsons–Zobel plots: An independent way to determine surface


complexation parameters?
Johannes Lützenkirchen ∗
Institut für Nukleare Entsorgung, Forschungszentrum Karlsruhe, Postfach 3640, D-76021 Karlsruhe, Germany
Received 1 June 2006; accepted 18 July 2006
Available online 24 August 2006

Abstract
Parsons–Zobel plots can in principle be used to estimate inner Helmholtz-layer capacitance values and electrochemical surface areas for
mineral particles. Their application to aqueous suspensions of various minerals has been documented in the literature. For the experimental data
used so far, the expected linear relationship between the overall and the diffuse-layer capacitances has been reported. The extracted values either
have not been used at all subsequently in a surface complexation model to describe the mineral surface charge versus pH curves, or were found
not to be suitable entirely for such purposes. In the latter case, the reported failure was not explained. In one part of the present paper, the Parsons–
Zobel plot concept is tested with data generated from a surface complexation model, for which the interfacial structure closely corresponds to
that assumed in the application of the Parsons–Zobel plot. From the analysis of the results it turns out that electrolyte binding and non-Nernstian
surface potential–pH curves more or less strongly affect the outcome of Parsons–Zobel plots. Despite the fact that the analysis in this paper is
restricted to iron(III) minerals only, it is concluded, in general, that the use of Parsons–Zobel plots with aqueous mineral suspensions to determine
inner Helmholtz-layer capacitances for subsequent application to surface complexation models cannot be recommended, since the reasons for
failure can be traced very nicely with applications to model-generated data. Such application requires the determination of further parameters,
and it was found that low electrolyte binding and Nernstian slopes should be imposed. Of these two issues, the more important is electrolyte
binding. For the surface complexation models, an inner Helmholtz-layer capacitance and weak electrolyte binding were required for a good fit to
experimental data. The values of the electrolyte binding constants required to achieve this end are in conflict with the assumptions of the Parsons–
Zobel plot (absence of specific adsorption). However, these parameters would not necessarily cause specific adsorption in terms of a classical
colloid chemistry definition (i.e., would not shift isoelectric points). The electrochemical surface areas were found to be in good agreement with
the value used to generate the data. Based on this, there is a potential for using the approach to determine surface areas in situ from titration curves.
Consequently, in a second part of the paper, Parsons–Zobel plots are applied to experimental data with the objective of determining electrochemical
surface areas in situ. Application to various sets of published experimental titration data for hydrous ferric oxide yielded consistently very large
electrochemical surface areas for fresh samples. This can be explained by very small particles and/or inclusion of substantial amounts of water in
the suspended particles. As would be expected, the electrochemical surface area for aged ferrihydrite was found to be substantially lower.
© 2006 Elsevier Inc. All rights reserved.

Keywords: Surface charge; Mineral suspensions; Surface complexation model; Parsons–Zobel plot; Specific adsorption; Nonspecific adsorption; Electrolyte
binding; Specific surface area

1. Introduction pacitance measurements, which are possible in such electrode


systems, can be directly applied in the use of Parsons–Zobel
The Parsons–Zobel plot was originally applied to electrode plots. This is usually not possible with mineral suspensions,
systems [1] to determine capacitance values and to evaluate which are the focus of the present work. As indicated above,
whether specific adsorption occurs at the electrode or not. Ca- the approach of Parsons and Zobel was originally advocated
as a means to detect “specific” adsorption of (electrolyte) ions.
Subsequently, the approach has been applied with the neces-
* Fax: +49 (0) 7247 82 3927. sary adjustments to suspensions of oxide minerals [2,3]. Here,
E-mail address: johannes@ine.fzk.de (J. Lützenkirchen). the purpose was to determine the (electrochemical) surface area
0021-9797/$ – see front matter © 2006 Elsevier Inc. All rights reserved.
doi:10.1016/j.jcis.2006.07.037
J. Lützenkirchen / Journal of Colloid and Interface Science 303 (2006) 214–223 215

of the minerals and the inner layer capacitance of the min- the need to fit titration data and to involve parameter interde-
eral/electrolyte interface. From the point of view of surface pendencies of the capacitance with site density and/or stability
complexation modeling, additional information on in situ elec- constants [7].
trochemical surface areas and a priori knowledge of specific As a consequence, the Parsons–Zobel plot would allow con-
capacitance values pertaining to a simple but useful picture sistent determination of the capacitance value and an in situ
of the mineral/electrolyte interface [2–4] would be very help- value for the specific surface area. These two parameters would
ful. not be affected by the site density parameter, and parameter in-
Specific surface area is important in such models, since, to- terdependencies would be avoided. This would clearly be an
gether with the site density of surface functional groups, it will advantage, since at present the site density parameter cannot
determine the available surface sites in a given adsorption ex- be experimentally determined [8]. Even if there are approaches
periment. In situ determined electrochemical surface areas may to calculating and interpreting capacitance values for the triple
be much more relevant to surface complexation studies than layer model, the inner layer capacitance currently largely re-
surface areas obtained from gas adsorption studies on dry pow- mains an adjustable (that is to say a fitted) parameter [9].
ders. Values for the latter may include certain problems, some The Parsons–Zobel plot has been previously applied to min-
of which are summarized below: eral suspensions and was discussed in some detail by Trasatti
and co-workers [2,3]. Part of the discussions concerned particle
– there is no unique standard method for the gas adsorption geometry, and the authors considered it necessary to introduce
procedure: the effect of particle geometry in the diffuse-layer capacitance.
• different commercial or lab-built setups exist (e.g., one- They did not do likewise for the inner layer capacitance, which
point, three-point, or full isotherm), which potentially is more difficult, since the thickness of the inner layer needs to
yield different values of the specific surface area and be known. Another part of the discussion concerned the Nerns-
generally cause a variation of data quality; tian behavior of the oxides, and the authors concluded that non-
• different probe molecules exist, which allow a more or Nernstian behavior would not affect the estimated electrochem-
less dense coverage of the surface, yielding (potentially ical surface areas. The Parsons–Zobel plot has been recently
depending on issues like surface roughness) in different re-discussed by Gunnarsson [4], and the conclusion was that its
values for the specific surface area of identical samples; application with experimental data to obtain parameters for sur-
• even with one probe molecule such as nitrogen (as the face complexation models failed to yield satisfactory results for
most frequently used probe), different orientations of the several systems. For example the extracted Parsons–Zobel sur-
molecule may be assumed, resulting in different specific face areas were much higher than the measured BET surface
surface areas for one set of raw data; areas for hematite (factor of about 1.5, background electrolyte
– there is a need to degas the samples at higher temperatures: NaNO3 ), magnetite (factor of about 4, background electrolyte
• degassing for different times and at different tempera- KNO3 ), and titanium oxide (factor of about 1.5, background
tures may influence results; electrolyte KNO3 ). An explanation for the high electrochemi-
• exposure of minerals to higher temperatures may affect cal surface areas was based on problems with BET analysis on
their state to some extent; dried particles as outlined above, and on the application of the
– there is a danger of particles in the dry state being inacces- planar geometry and potential coagulation effects, in particular
sible to gas molecules: at high ionic strength. As for the use of surface complexation
• particles may stick together and part of the surface may models, it is noteworthy that Gunnarsson [4] has used planar
not be accessible to the probe molecules, e.g., gibbsite, geometry without electrolyte binding. This led to a good agree-
where a huge discrepancy has been reported between gas ment with data on goethite in NaClO4 as discussed in more
adsorption results and AFM imaging [5]; detail below. For hematite in NaNO3 the approach failed to
• particles may stick to walls and part of the surface may give an acceptable description of the surface charge data. It was
not be accessible to the probe molecules. reported that inclusion of electrolyte binding and a significant
decrease of the capacitance was necessary to obtain a good fit
This summary justifies the question of whether an in situ to the data. In some of the other data sets, it can be seen that
method would not be most relevant to any subsequent work a good description was not possible with straightforward appli-
with suspensions. The EGME method [6] is such an approach. cation of Parsons–Zobel plot results either. This failure was not
The Parsons–Zobel plot concept would be an alternative way to fully explained.
estimate suspension surface area in situ directly from titration In the present paper the Parsons–Zobel plot is applied to
data, with the benefit of simultaneously yielding the specific ca- model-generated data. For these, all surface complexation pa-
pacitance value of the Stern layer. rameter values are known and the ability of the approach to
Specific capacitance values are relevant to all surface com- retrieve the known parameter values can be tested. One as-
plexation models involving layers different from the diffuse sumption of Trasatti and co-workers [2,3] was that Nernstian
layer of mobile counterions. Thus the most realistic surface behavior of the mineral occurs in the vicinity of the point of zero
complexation models are concerned. In particular, the basic charge. This assumption can easily be tested with the gener-
Stern model (BSM) includes only one such layer, and its capaci- ated data, since all potentials are obtained from the calculations.
tance could be directly inferred from experimental data without Since the electrostatic model concept used in the calculations
216 J. Lützenkirchen / Journal of Colloid and Interface Science 303 (2006) 214–223

to generate the data (i.e., the BSM) is in principle consistent interfacial potentials via
with the one assumed in the Parsons–Zobel plot concept, a con-
clusion about the actual compatibility of the interfacial models C1 = σ0 /(Ψ0 − Ψ1 ). (4)
should be obtained. After the application to generated data, the In Eq. (4), Ψi is the interfacial potential in the adsorption
Parsons–Zobel plot is applied to experimental surface charge plane i. Plane “0” is the surface plane, whereas plane “1” corre-
data for hydrous ferric oxide (HFO). This system was chosen sponds to the inner Helmholtz plane (IHP). For the basic Stern
because some speculation about the specific surface areas of model the potential at the IHP corresponds to the diffuse-layer
HFO can be found in the literature. The specific surface areas potential; i.e., Ψd = Ψ1 .
typically measured on dry samples are about 250 m2 /g. How- The application of Parsons–Zobel plots requires values for
ever, in most modeling studies the value is set to 600 m2 /g the overall capacitance and the capacitance of the diffuse layer.
based on a recommendation by Dzombak and Morel [10]. The The overall (mass-specific) capacitance is obtained from the
system is considered as a good example of the disagreement first derivative of the mass-specific surface charge vs pH curve
between gas adsorption measurements in the dried state and ac- at the pristine point of zero charge, (∂σ0 /∂pH)ppzc , in the fol-
tual surface area in suspension and was, therefore, chosen here lowing fashion:
to evaluate the Parsons–Zobel plot approach.
Ctot,M = (∂σ0,M /∂Ψ0 )ppzc
2. Theory = (∂σ0,M /∂pH)ppzc (∂pH/∂Ψ0 )ppzc . (5)
(∂σ0,M /∂pH)ppzc can be obtained from titration data. For
The overall (experimental) capacitance, Ctot,A in F/m2
(∂pH/∂Ψ0 )ppzc in the original applications of Parsons–Zobel
(A indicates surface area), within the basic Stern model is given
plots to oxide suspensions, it was assumed that Nernstian be-
by the equation
havior of the surface can be used. This has been pursued by
1/Ctot,A = 1/C1,A + 1/Cd , (1) Gunnarsson [4] based on work by Larson and Attard [11]. How-
ever, it is also possible to have another point of view based on
with the first r.h.s. term involving the Stern layer capacitance part of the available experimental data on oxides [12]. Over-
and the second the diffuse-layer capacitance. The above equa- all, there is no agreement on the slope of surface potential pH
tion can be related to the electrochemical specific surface area curves for oxide minerals, and therefore this is an inherent un-
(S in m2 /g) as follows: certainty in the application of the Parsons–Zobel plot to oxide
minerals.
S/Ctot,M = S/C1,M + 1/Cd . (2)
The subscript M indicates that the overall and Stern layer ca- 3. Data collection and data treatment
pacitances are on a mass-specific basis, in F/g. The (surface-
specific) diffuse-layer capacitance at 298.15 K can be calcu- 3.1. Generated data
lated from
Model data were generated with the speciation code ECO-
Cd = εr ε0 κ (3) SAT [13] based on goethite surface complexation parameters
given by Rietra et al. [14] for a range of sodium nitrate con-
with εr the dielectric constant of water, ε0 the permittivity of centrations. Sodium nitrate was chosen because it is has been
free space, and κ the inverse Debye length. This equation is de- shown that it does not shift isoelectric points of goethite (i.e.,
rived from the Gouy–Chapman equation either as the integral sodium nitrate is a nonspecifically adsorbing background elec-
or differential capacitance of the diffuse layer for sufficiently trolyte, based on the common colloid chemistry criterion).
low surface potentials at the head end of the diffuse layer (i.e., A mass-specific surface area of 100 m2 /g is a typical value for
<26 mV). The condition of sufficiently low surface potential this material, when it is prepared by slow titration of an iron(III)
holds in particular for the isoelectric point (IEP), which is in solution, and this value was used in all calculations presented
general close to the point of zero surface charge. For aque- here.
ous solutions
 at 25 ◦ C, Eq. (3) results in Cd (IEP) (F m−2 ) = For all sodium nitrate concentrations considered, the output
2.285 Ic (mol dm−3 ) = 7 × 10−10 · κ (m−1 ), where Ic is the from ECOSAT was used to calculate proton-associated (i.e.,
ionic strength and κ the inverse Debye length. Note that the measurable) mass-specific surface charge in C/g as a func-
diffuse-layer capacitance is always per area, whereas the Stern tion of pH. Furthermore, the potential in the surface plane was
layer and the overall capacitances may be per area or per mass. retrieved from the output and plotted as a function of pH to
Among the electrostatic interfacial concepts of the common verify the assumption of Nernstian behavior inherent in the ap-
surface complexation models, the one that most closely corre- plication of Parsons–Zobel plots to mineral suspensions [2,3].
sponds to that of the Parsons–Zobel plot approach is the basic For a given sodium nitrate concentration, mass-specific surface
Stern model. That is why recently it was attempted to link charge (in the surface plane) as a function of pH was spline fit
Parsons–Zobel plot results to this particular model. Within this and the minimum derivative was used to calculate Ctot,M from
model C1 is the capacitance of the Stern layer and is related Eq. (5) (assuming Nernstian behavior in the very first step).
to the measurable surface charge density, σ0 , and the unknown The results for the various ionic strengths were then plotted as
J. Lützenkirchen / Journal of Colloid and Interface Science 303 (2006) 214–223 217

1/Ctot,M vs 1/Cd , calculated from Eq. (3). From linear regres- of zero charge (results shown in Table 1). Analysis over the
sion S and C1,M were obtained; cf. Eq. (2). It was observed that whole pH range yielded values with lower slopes. From the re-
the minimum derivative of the charging curve coincided with sults in Table 1 it becomes obvious that Nernstian behavior is
the point of zero charge. Recently, it has been suggested to use approached when the ionic strength is lowered. However, true
derivatives of charging curves to determine the point of zero Nernstian behavior was never reached, even with unrealistically
charge [15], a procedure that is confirmed with generated data low values of ionic strength. The deviation between assumed
here. Subsequently, the “true” model inherent surface potential and actual behavior of the surface potential is an aspect already
pH curves were used in the evaluations to replace the Nernst discussed by Trasatti and co-workers [2,3] and is addressed here
assumption in Eq. (5). with the objective of evaluating the applicability of Parsons–
To understand the importance of Nernstian behavior and of Zobel plots for BSM parameterization.
electrolyte binding in a second step, the following parameter The Parsons–Zobel plot for the model-generated data using
variations were carried out, while keeping all other parameters the complete set of parameters from Rietra et al. [14] for sodium
constant: nitrate is shown in Fig. 1 and the outcome is summarized in
Table 2. This is compared to curves that would result directly
– variation of the electrolyte binding constant at the standard from the known input parameters using Eqs. (1) and (3). The
capacitance value; parameters derived from the treatment of the model-generated
– variation of the capacitance value at the standard electrolyte data under the assumption of Nernstian behavior are 1.12 F/m2
binding constant. for the capacitance of the Stern layer (compared to the value of
0.91 F/m2 used to generate the data) and about 104 m2 /g for
Increasing the electrolyte binding constant and the model ca- the specific surface area (compared to the value of 100 m2 /g
pacitance value resulted in steeper charging curves and more used to generate the data); cf. Fig. 1a. This could not be im-
pronounced non-Nernstian behavior. proved by using the actual slopes of surface potential pH curves
(from Table 1); cf. Fig. 1b. Thus deviation from Nernstian be-
3.2. Experimental data havior is not the reason for the obtained discrepancy in this
case. As noted by Trasatti and co-workers [2,3], deviation from
Different sets of HFO data were taken from Cox and the ideal Nernst behavior is not expected to result in errors in
Gosh [16], Schwertmann and Fechtner [17], Swallow [18], the estimated “electrochemical” surface area. Indeed, with the
Hansen et al. [19], and Trivedi et al. [20]. They were trans- generated data and a Nernst slope, a value close to the input
formed to mass-specific surface charge vs pH curves and spline value is obtained (Table 2). Use of a different slope will just
fit. The minimum of the derivative was used to compute the parallel-shift the Parsons–Zobel plot and therefore not affect
overall mass-specific capacitance assuming Nernstian behav- the extracted electrochemical surface area (which depends on
ior. Its inverse was plotted vs 1/Cd as calculated from Eq. (3). the slope of the Parsons–Zobel plot only). However, an aspect
Linear regression yielded the values for the mass-specific inner not discussed by Trasatti and co-workers [2,3] is the inherent
layer capacitance and the electrochemical specific surface area. problem caused if the surface potential–pH slopes not only de-
viate from Nernstian behavior but also vary with ionic strength
4. Results and discussion (cf. Table 1). This will actually result in problems in estimat-
ing the electrochemical surface area and may explain the small
4.1. Applications to generated data deviations obtained with the generated data. The deviations are
small since the changes in the slope of the potential–pH curves
The derivatives of the surface potential versus pH curves for are small. If the variation of the slopes with ionic strength were
the model parameters according to Rietra et al. [14] are sum- to become larger, the error in the estimated electrochemical sur-
marized in Table 1. It can be noted that the slope of the surface face area based on a constant Nernst slope would become more
potential versus pH curves became more Nernstian when the important. Differences in activity coefficients at the surface and
range of data was limited to values in the vicinity of the point in the bulk solution (apart from issues related to classical elec-
trostatic terms) may contribute to deviations from Nernstian
Table 1
behavior and to the variations with ionic strength.
Slope of surface potential versus pH curves for the goethite model of Rietra et
al. [14] The electrochemical surface area furthermore pertains to the
Gouy–Chapman equation. It is referred to the same mass as the
I in M Slope in mV
BET surface area, but due to the separation of the head end of
0.1 56.71
0.01 58.00
the diffuse layer, the electrochemical surface area should dif-
0.002 58.58 fer from the zero-plane surface area. It should be larger if the
0.001 58.73 zero-plane area is ideal. In the case of the flat-plane geometry of
6 × 10−4 58.82 the particles assumed here, this difference should not show up.
1 × 10−4 59.00 Though it might be of interest to investigate the issue in more
2 × 10−5 59.07 detail for ideal spherical or cylindrical geometries, the diffuse-
1 × 10−5 59.08
layer charge–potential relationship will usually simplify to the
2 × 10−6 59.09
limiting relationship for the differential capacitance of the dou-
218 J. Lützenkirchen / Journal of Colloid and Interface Science 303 (2006) 214–223

(a)

(b)

Fig. 1. Parsons–Zobel plot for the goethite system (generated data based on parameters by Rietra et al. [14] for a basic Stern model; original parameters: 0.91 F/m2 ,
100 m2 /g). (a) The Parson–Zobel plot based on Nernst behavior. (b) The actual potential–pH relationship as obtained from the data generation. The drawn lines and
triangles show values that would be directly calculated using the original parameters.

Table 2
Results of the application of Parsons–Zobel plots to the different parameterizations of surface complexation models
Comments on model parameters Nernst slope Actual slope
APZ in m2 /g C in F/m2 APZ in m2 /g C in F/m2
Original surface complexation parameters 104.3 1.12 104.4 1.25
Electrolyte binding constants decreased by two orders of magnitude 100.2 0.83 100.2 0.91
Electrolyte binding constants decreased by four orders of magnitude 100.1 0.83 100.1 0.91
Note. Original surface complexation parameters: 0.91 F/m2 , 100 m2 /g; electrolyte binding constant log K = −1; parameters from Rietra et al. [14] within a basic
Stern model. Variants reported here include the same model parameters except for electrolyte binding constants set to log K = −3 and −5, respectively.

ble layer, i.e., Eq. (3). So for theoretical data sets generated by Another problem may occur in comparing results based on
model parameters, no real changes are expected. For actual par- Eq. (5) with a surface complexation model including Eq. (4).
ticles, however, discrepancies between measured surface areas While the former uses a differential capacitance, the latter is an
and electrochemical surface areas may be explained by the non- integral capacitance. Derivation of σ0 with respect to pH (based
ideal geometries of the real particles. on Eq. (4)) yields the following expression (the suffix ppzc is
J. Lützenkirchen / Journal of Colloid and Interface Science 303 (2006) 214–223 219

omitted for simplicity in the following) for the surface specific Based on Eq. (18), Eq. (10) only holds if the charge density
capacitance: in plane 1 is zero (or sufficiently close to zero). This condition
will be approximately fulfilled for weak ion binding or at suf-
C1 = (∂σ0 /∂Ψ0 )/(1 − ∂Ψd /∂Ψ0 ). (6) ficiently low ionic strength. In other words, the Parsons–Zobel
Insertion into Eq. (1) and using Eq. (5) yield plot is not applicable to delivering BSM surface complexation
parameters in the case of too strong electrolyte binding.
1/Ctot = 1/(∂σ0 /∂Ψ0 ) This result is corroborated by model parameter variation
 
= (1 − ∂Ψd /∂Ψ0 )/(∂σ0 /∂Ψ0 ) + 1 −(∂σd /∂Ψd ) , (7) and subsequent application of Parsons–Zobel plots, with the
goethite model as a reference system. Variation of the (sym-
and rearranging results in metrical) electrolyte binding constants from log K = −1 to
−(∂σd /∂Ψd ) = −(1 − ∂Ψd /∂Ψ0 ) × (∂σd /∂Ψd ) −5 for otherwise constant parameters showed that the consis-
tency between input parameters and Parsons–Zobel plot para-
+ (∂σ0 /∂Ψ0 ) (8)
meters indeed improves. Excellent agreement is obtained for
and finally log K  −3 (Table 2). Using data-derived (actual) surface po-
tential versus pH slopes instead of Nernstian slopes does not
(∂σd /∂Ψ0 ) = −(∂σ0 /∂Ψ0 ) (9) contribute to improving the results at high electrolyte binding
or constants, whereas with lower electrolyte binding constants the
use of the actual potential–pH relationship leads to the correct
(∂σd /∂σ0 ) = −1. (10) capacitance value. Variation of the inner layer capacitance at
In these equations σd is the diffuse-layer charge. Equa- constant electrolyte binding corroborates this conclusion: al-
tion (10) must hold for the original Parsons–Zobel plot to be though a decrease of the capacitance makes the surface more
applicable in the context of the tested surface complexation Nernstian, it is found that this contribution to improving the ac-
model. Inclusion of electrolyte binding (in this kind of model curacy of the Parsons–Zobel plot parameters is not as important
usually in plane 1) yields the following equation based on over- as the contribution due to counterion binding (Table 2).
all particle charge neutrality: Electrolyte binding appears to be rather strong, as indicated
by shifts of IEPs at higher ionic strength or unsymmetrical
σd = −(σ0 + σ1 ), (11) charging curves. Shifts of IEPs at higher ionic strength appear
where σ1 is the charge density in the plane of electrolyte ad- to be a very common feature [21,22].
sorption. Corresponding to Eq. (10), differentiation of Eq. (11) The results may also be used to explain certain observa-
with respect to σ0 yields tions made by Gunnarsson [4]. In that work it was found that
Parsons–Zobel plots would yield parameters that could be used
(∂σd /∂σ0 ) = −1 − (∂σ1 /∂σ0 ). (12) in a straightforward manner to describe charging of goethite in
NaClO4 solutions (note in this context that in the BSM model
Combining Eqs. (10) and (12) yields
by Gunnarsson [4] no electrolyte binding was involved at all!).
(∂σ1 /∂σ0 ) = 0. (13) The explanation of the discrepancy between the outcome of
Parsons–Zobel plots with goethite data used in this work and
The derivative ∂σ1 /∂σ0 within the basic Stern model becomes those used by Gunnarsson [4] is that Gunnarsson’s titrations
more transparent by using the surface species in the planes of in NaClO4 solutions produced surface charge densities that are
adsorption; i.e., much lower than those corresponding to Rietra’s model in NaCl
   
σ1 = −B × SOH2 · · ·A– – SOH−1/2 · · ·C+ , or NaNO3 . In other words, the electrolyte binding in NaClO4
1/2
(14)
 1/2 
solutions is of minor importance. From Rietra’s titrations in var-
σ0 = B × SOH2 ious electrolytes it is known that the surface charge density is
     
+ SOH2 · · ·A– – SOH−1/2 – SOH−1/2 · · ·C+ lower in NaClO4 solutions than in NaCl or NaNO3 solutions.
1/2
 1/2    This was also discussed by Gunnarsson [4]. The results of the
= B × SOH2 – OH−1/2 − σ1 . (15) present work may also be used as an explanation for the failure
Combination of the two yields of the Parsons–Zobel plots to yield a description of the hematite
 data in NaNO3 used by Gunnarsson [4]. Gunnarsson indicated
1/2   
σ1 = B × SOH2 – SOH−1/2 − σ0 , (16) that involvement of electrolyte binding would improve the fit
with a simultaneous decrease of the capacitance value. This
where B is a factor relating molar concentrations of surface
corresponds directly to the parameter variation results obtained
species to surface charge densities. Equation (13) requires that
with goethite in this work. So the interpretation of the failure
  1/2    
(∂σ1 /∂σ0 ) = B × ∂ SOH2 – OH−1/2 ∂σ0 − 1 based on the results of this paper would be that electrolyte bind-
ing is too strong in the hematite system studied by Gunnarsson
=0 (17)
[4] for a Parsons–Zobel plot to be applicable, whereas Gunnars-
and finally son explained the failure by coagulation effects of the hematite
  1/2     during titration. Coagulation may be related to the strong ad-
B × ∂ SOH2 – SOH−1/2 ∂σ0 = 1. (18) sorption of the electrolyte ions.
220 J. Lützenkirchen / Journal of Colloid and Interface Science 303 (2006) 214–223

When Parsons–Zobel plots are to be used in parameteriza- ionic strength >0.1 M. These were therefore omitted from the
tion procedures for surface complexation models, the present discussion.
results suggest that an optimum parameterization might be a In general, the Parsons–Zobel plots yielded very high spe-
sufficiently high capacitance value (to grant the observed pro- cific surface areas of 2000 m2 /g and more. A notable exception
ton uptake) combined with sufficiently low electrolyte binding. is obtained with the data by Hansen et al. [19] with a value
Parsons and Zobel [1] implied that their methodology as ap- of around 500 m2 /g. This may be explained by the fact that
plied to the mercury electrode holds in the presence of non- Hansen et al. aged their sample 9 days, whereas the other sam-
specifically adsorbing ions. The same holds for mineral sus- ples may rather represent fresh ferrihydrite. The particle size
pensions. This has not received any attention in previous ap- of those latter samples must be extremely small or the den-
plications of Parsons–Zobel plots to mineral suspensions. The sity must be substantially affected by inclusion of water, if
term specifically or nonspecifically bound may unfortunately the results from the Parsons–Zobel plots are to be meaningful.
lead to confusion. Traditionally, the terminology with mineral The “electrochemical” surface area may also significantly differ
suspensions has involved an operational definition, according from BET or EGME areas, since it refers to the Gouy–Chapman
to which specifically binding electrolyte ions are able to shift equation. In general the particles will be far from having a per-
points of zero charge. From electroacoustic measurements on fect geometry, but they will instead exhibit surface roughness.
various oxide minerals at high ionic strength with electrolytes, To what extent this interferes with the Parsons–Zobel plot as-
which are “inert” at lower ionic strength, it is known that iso- sumptions cannot be discussed based on the simple models used
electric points shift [21,22]. Thus in terms of this definition it here, but it would be an interesting issue for theoretical consid-
might be sufficient to go up in ionic strength to show that the erations. One might expect that the head end of the diffuse layer
monovalent ions are never “inert” over the whole concentration would assume a more regular shape than an underlying rough
range. surface. It is clear that surface roughness creates a higher area
Parsons and Zobel [1] indicate for their electrode systems compared to a mean surface at a smooth model surface “zero”
that a linear Parsons–Zobel plot is an indication of nonspe- plane. The separation of the location of the “electrochemical”
cific binding. This clearly does not hold for mineral suspensions
surface area from the surface would result in a higher value
based on the presented calculations, since linear plots were al-
for the “electrochemical” surface area, compared to the surface
ways obtained in the present examples. However, parameter
area for the smooth model surface “zero” plane. Since all val-
variation showed that with sufficiently strong electrolyte bind-
ues refer to the same mass of sorbent, it is the interplay between
ing constants, the original parameters could not be retrieved.
actual surface roughness and the extent of the separation and
4.2. Applications to experimental data the actual size of the particles, which will determine the ratio
between the surface area of the rough surface and the “elec-
As has been seen in the treatment of generated data, trochemical” surface area. As for the probe molecules used in
Parsons–Zobel plots yield good estimates for specific surface BET measurements the dimensionality of the surface roughness
areas, even if the slopes of surface potential–pH curves vary and the size of the probe molecules come into play. As a con-
with ionic strength. Thus, it was attempted to use the approach sequence, it is not possible to give a general tendency, although
with hydrous ferric oxide (HFO). Literature data [16–20] were one would expect that the roughness effect might dominate. But
used to this end. The results are summarized in Table 3, in- this is then a theoretical expectation, since the comparison must
cluding the extracted capacitance and electrochemical surface presently be made through measurement with probe molecules
area values as well as the number of points used for the regres- of finite size. The model involved in the interpretation always
sion and the linear regression coefficient. The linear regressions assumes ideal geometries, and this may lead to inconsistencies
were usually very good. Interestingly for the data by Swallow between measured values and model-input values.
[18] no extreme value in the derivative was obtained for the From the application of the Parsons–Zobel plot it is ob-
highest ionic strength, a result that would be expected since the served that the extracted capacitance values vary substantially.
charging curves typically assume a linear pH-dependency at The area specific capacitance depends on the slope of the mass-
sufficiently large electrolyte concentrations. The same was ob- specific surface charge versus pH curve around the pH of the
tained for a number of other data sets, which included data at slope extremum and the extracted surface area. The combina-

Table 3
A summary of results of the application of Parsons–Zobel plots to various HFO samples
Sample Electrolyte S in C in Range of pH for Number of points Linear regression
from in titration m2 /g F/m2 slope extremum used in regression coefficient
[16] NaNO3 2000 0.31 8.15–8.38 3 1.000
[17] KNO3 2690 1.03 8.13–8.14 3 0.998
[18] NaClO4 1800 0.50 8.32–8.41 2a NA
[19] NaClO4 525 0.18 6.71–7.00 3 0.979
[20] NaNO3 2650 0.12 6.54–7.44 3 1.000
a The derivative at the highest ionic strength did not yield an extremum for this data set. The linear regression coefficient is therefore meaningless and the results
are consequently to be considered with care.
J. Lützenkirchen / Journal of Colloid and Interface Science 303 (2006) 214–223 221

Fig. 2. Comparison of the mass-specific surface charge of the HFO samples that were analyzed by Parsons–Zobel plots. Data from Schwertmann and Fechtner [17],
Cox and Gosh [16], Swallow [18], Hansen et al. [19], and Trivedi et al. [20].

tion of charging curves with a low slope and a high area results tablished, the maxima of the derivatives of the more aged sam-
in low capacitance values. This is not necessarily seen from ples [19,20] are significantly lower than pH 8. Furthermore, for
Fig. 2, where mass-specific surface charge versus pH curves are the data by Sparks and co-workers [20], they vary over nearly
shown for the different samples. The data by Hansen et al. [19] one pH unit as a function of ionic strength. Nevertheless, in
clearly show the lowest charge, which may be attributed to the the respective papers the point of zero charge is fixed at pH 8.
extended aging time of their solid. The aging time for the sam- This is done based on other assumptions/hypotheses (e.g., com-
ple by Sparks and co-workers [20] was two days, which is also mon intersection point). Rigorous determination of points of
longer than what is reported for the fresh samples prepared by zero charge from titrations is difficult, since common intersec-
Cox and Gosh [16] and Swallow [18]. Finally, it is interest- tion points are also obtained in 2:1 electrolytes and are different
ing to note from Fig. 2 that apparently the data for the fresh from those obtained in “nonspecific” electrolytes. Thus it is
preparations [17,18] and those for the two slightly aged sam- highly preferable to apply some electrokinetic method to cor-
ples [16,20] agree with each other but not with the other set of roborate the point of zero charge, in particular at low ionic
data. The exact aging for the sample from Schwertmann and strength, where the resolution of microelectrophoresis increases
Fechtner [17] is not known, but it was reported that the sample and electrolyte specific effects are minimized. The problem is
was freeze-dried. The salts used as background electrolytes in the low solid concentration in such experiments, which can be
the titrations (cf. Table 3) do not seem to play a significant role, avoided in electroacoustic measurements. An ill-defined point
although they usually do for goethite at 0.1 M electrolyte (i.e., of zero charge will strongly affect the parameterization of a
charging is stronger in nitrate than in perchlorate media [14]). surface complexation model (e.g., in the case of Ref. [19] the
Sparks and co-workers [20] actually discuss the discrepancy be- adsorption modeling for silicic acid and in the case of Ref. [20]
tween the amount of charge on their samples and that on other the adsorption modeling for lead), if the postulated metal ion or
HFO samples. They also mention the difference of surface spe- ligand surface species is charged. Net transfer of charge to or
cific charge density between HFO and goethite. This difference from the inner-surface plane will be affected by the electrosta-
would dramatically decrease if the electrochemical surface ar- tic term involving the surface potential, which depends on the
eas obtained in this study from the Parsons–Zobel plots were smeared-out surface charge, and this effect can easily amount
applied, i.e., if HFO surface areas increased by a factor of 3 to orders of magnitude.
to 4. The capacitance values obtained are in all cases (except one)
Such a huge difference between electrochemical and BET quite low. From the tests on generated data, it is clear that these
surface areas has also been found by Gunnarsson [4], in partic- values are not expected to be useful. The very low values would
ular for magnetite, and this was explained by sticking together imply that the thickness of the inner Helmholtz layer is rather
of small particles after drying (see Introduction). large, that the relative permittivity of that layer is rather small
Another interesting issue in the treatment of the HFO data is (i.e., strong structuring of water), or both.
that the extreme values of the derivatives varied more strongly
than expected with respect to their position on the pH scale. 5. Conclusions
Bourikas et al. [15] have suggested that this extreme should
correspond to the point of zero charge. Whereas a point of zero Parsons–Zobel plots have been tested on model-generated
charge around pH 8 for HFO [10] seems surprisingly well es- data based on surface complexation parameters for goethite
222 J. Lützenkirchen / Journal of Colloid and Interface Science 303 (2006) 214–223

(within the BSM) and on experimental data on hydrous fer- The estimates for specific surface areas obtained from the
ric oxide (HFO). The application of the Parsons–Zobel plot to application of Parsons–Zobel plots to experimental data for
model-generated data led to the conclusion that two issues may HFO has resulted in unexpectedly high values. This agrees with
prohibit the use of parameters extracted from this approach in findings by Gunnarsson [4]. With the generated data, the es-
surface complexation models: timated specific surface areas were also higher (only slightly,
though) than the input model parameter values, but those re-
– Systems with significant electrolyte binding should not be sults suggested that quite reasonable agreement with expected
treated by the Parsons–Zobel-plot approach. In the present values can be obtained. The major obstacle here was found to
case, only if electrolyte binding occurred with log K  −3, be a variation of the surface potential vs pH curves with ionic
Parsons–Zobel plots were able to reproduce the expected strength.
parameters. For larger electrolyte binding constants, the Based on the reported results, at present it seems that the use
ion pairing interferes too strongly with the assumptions in- of the conventional Parsons–Zobel plots is not a general means
herent in the Parsons–Zobel-plot approach. This has been of independently estimating parameters. Although the present
shown with the generic example based on parameter varia- work is restricted to iron(III) minerals only, it is concluded from
tion but became also evident from underlying equations. the results that the approach can be used, but entails certain
– Nernstian slopes for surface potential versus pH curves of ranges for remaining unknown parameters (which need to be
relevant oxide minerals have been postulated but there is fitted to experimental data) and in a strict sense also for the
no ultimate proof for their governing the oxide/water sys- surface potential–pH curves. The approach yielded reasonable
tems. As an example in the model-generated data, the ac- values for the electrochemical surface area with the generated
tual slopes were not Nernstian and they also varied with data, but unexpectedly high values for most HFO samples com-
ionic strength. Decreasing strength of electrolyte binding pared to recommended literature values [10]. Although this
tended to create slopes closer to the Nernstian one. Note might be explained by assuming very small particles in com-
that the calculations to generate the data involve model as- bination with a low particle density and although it would place
sumptions, which await proof. However, the parameter set the surface charge density of HFO in a range similar to that of
by Rietra et al. [14] describes experimental surface charge goethite, the values for electrochemical surface areas obtained
very well for a wide range of conditions. by Parsons–Zobel plots have to be considered with caution,
one obstacle being that the slope of surface potential versus
The deviation from Nernstian behavior was found to be the pH curves is usually not accessible. It might be expected that
less important issue. the recent approach by Kallay et al. [29] might resolve this is-
When Parsons–Zobel plots are used in the context of para- sue. Surface potential–pH curves are measurable (single-crystal
meter estimation, it should subsequently be attempted to im- electrodes) by that approach, so that from σ0 (pH) experimental
pose sufficiently low electrolyte binding and Nernstian slopes. functions (potentiometric acid–base titrations of concentrated
This is the only way to assure consistency between the sur- suspensions) one may experimentally determine dσ0 /dΨ0 .
face complexation model and the assumptions inherent to the Partial incompatibility of the Parsons–Zobel plot capaci-
Parsons–Zobel plots. The present study explains observations tances with surface complexation models, which was previ-
by Gunnarsson [4], which indicate that Parsons–Zobel plots do ously noted [4], could be explained in the present paper. It is
not generally work as independent approaches to determine pa- due to the fact that the assumption of nonspecific adsorption of
rameters (capacitance values) for the basic Stern model (even the electrolyte ions is a question of definition. In colloid chem-
without inclusion of electrolyte binding). Conceptionally, in istry the term is often connected to observations of shifting iso-
surface complexation models, such as the basic Stern and other electric points. This notion is not necessarily compatible with
multilayer models, electrolyte ions actually bind to surface sites the application of Parsons–Zobel plots. The electrolyte binding
as so-called outer-sphere surface complexes as point charges. constants commonly encountered in basic Stern or triple-layer
Sometimes, and particularly in the case of temperature stud- models are too high compared to the notion within the Parsons–
ies, rather strong electrolyte binding constants are required [23] Zobel plot.
to fit experimental data. Consequently, in its original version
the Parsons–Zobel plot will not be a realistic option to estimate Acknowledgments
capacitance values for the basic Stern model in general, par-
ticularly since it is known that electrolyte binding is related to The author is grateful for comments and suggestions from
the nature of the electrolyte and that of the surface [24–27]. Professor Kallay on a previous version of this manuscript prior
Structure-breaking and -making properties of the electrolyte to submission. The author also gratefully acknowledges the
ions and the surface with respect to the water molecules will af- comments by an anonymous reviewer, which were very help-
fect both the interlayer capacitance (via distance and dielectric ful in clarifying certain parts of the work.
constant) and potentially even the electrochemical surface area.
Attempts have recently been made to incorporate water struc- References
ture into surface complexation models [28]. These attempts
currently do not include the more elaborate, but qualitative ex- [1] R. Parsons, F.G.R. Zobel, J. Electroanal. Chem. 9 (1965) 333.
planations by Dumont et al. [24–27]. [2] A. Siviglia, A. Daghetti, S. Trasatti, Colloids Surf. 7 (1983) 15.
J. Lützenkirchen / Journal of Colloid and Interface Science 303 (2006) 214–223 223

[3] S. Ardizzone, S. Trasatti, J. Electroanal. Chem. 251 (1988) 409. [16] C.D. Cox, M.M. Gosh, Water Res. 28 (1994) 1181.
[4] M. Gunnarsson, Ph.D. thesis, Göteborg University, Göteborg, Sweden, [17] U. Schwertmann, H. Fechtner, Clay Miner. 17 (1982) 471.
2002. [18] K.C. Swallow, Ph.D. thesis, Massachusetts Institute of Technology, Cam-
[5] J. Rosenqvist, P. Persson, S. Sjöberg, Langmuir 18 (2002) 4598. bridge, MA, 1978.
[6] D.L. Carter, M.D. Heilmann, C.L. Gonzalez, Soil Sci. 100 (1965) 356. [19] H.C.B. Hansen, T.P. Wetche, K. Raulund Rassmussen, O.K. Borggaard,
[7] J. Lützenkirchen, J. Colloid Interface Sci. 210 (1999) 384. Clay Miner. 29 (1994) 341.
[8] J. Lützenkirchen, J.F. Boily, L. Lövgren, S. Sjöberg, Geochim. Cosmo- [20] P. Trivedi, J.A. Dye, D.L. Sparks, Environ. Sci. Technol. 37 (2003) 908.
chim. Acta 66 (2002) 3389. [21] M. Kosmulski, J.B. Rosenholm, Adv. Colloid Interface Sci. 112 (2004)
[9] D.A. Sverjensky, Geochim. Cosmochim. Acta 69 (2005) 225; 93.
D.A. Sverjensky, Geochim. Cosmochim. Acta 65 (2001) 3643. [22] W.N. Rowlands, R.W. O’Brien, J. Colloid Interface Sci. 175 (1995)
[10] D.A. Dzombak, F.M.M. Morel, Surface Complexation Modelling Hydrous 190.
Ferric Oxide, Wiley, New York, 1990. [23] M.L. Machesky, D.J. Wesolowski, D.A. Palmer, K. Ichiro-Hayashi,
[11] I. Larson, P. Attard, J. Colloid Interface Sci. 227 (2000) 152. J. Colloid Interface Sci. 200 (1998) 298.
[12] L. Bousse, N.F. De Rooij, P. Bergveld, IEEE Trans. Electron Devices 30 [24] F. Dumont, A. Watillon, Discuss. Faraday Soc. 52 (1971) 324.
(1983) 1263. [25] F. Dumont, D. Van Tan, A. Watillon, J. Colloid Interface Sci. 55 (1976)
[13] A. Keizer, W. van Riemsdijk, ECOSAT, vers. 4.7, Wageningen University, 678.
2001. [26] F. Dumont, J. Warlus, A. Watillon, J. Colloid Interface Sci. 138 (1990)
[14] R.P.J.J. Rietra, T. Hiermstra, W.H. van Riemsdijk, J. Colloid Interface 543.
Sci. 240 (2001) 384. [27] F. Dumont, S. Contreras, M. Diaz y Alonso, Anal. Quim. 91 (1995) 635.
[15] K. Bourikas, C. Kordulis, A. Lycourghiotis, Environ. Sci. Technol. 39 [28] T. Hiemstra, W.H. van Riemsdijk, J. Colloid Interface Sci. 301 (2006) 1.
(2005) 4100. [29] N. Kallay, Z. Dojnovic, A. Cop, J. Colloid Interface Sci. 286 (2005) 610.

You might also like