You are on page 1of 7

Applied Clay Science 87 (2014) 205211

Contents lists available at ScienceDirect

Applied Clay Science


journal homepage: www.elsevier.com/locate/clay

Research paper

Ion exchange in amorphous alkali-activated aluminosilicates:


Potassium based geopolymers
Taisiya Skorina
Department of Material Science and Engineering, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Room 4-217, Cambridge, MA 02139, USA

a r t i c l e

i n f o

Article history:
Received 4 October 2012
Received in revised form 30 September 2013
Accepted 2 November 2013
Available online 23 November 2013
Keywords:
Geopolymers
Ion exchange
Inorganic sorbent

a b s t r a c t
Ion exchange properties of pure potassium-based geopolymer (pure K-GP) formulated with Si/Al = 2.1 and
K/Al = 1.0 0.02 atomic ratios have been studied. For pure K-GP, the maximum ion-exchange level of K+
with respect to Na+ and Cs+ is 77 and 61%, respectively, as determined by exhausted exchange with 1 M solutions of appropriate nitrates at 22 2 C. The geopolymer shows a slight preference for K+ over Cs+ as indicated
by the normalized Cs/K-GP isotherm obtained under noncompetitive exchange. Parent geopolymer yields a pure
geopolymer when it is extensively washed with deionized water until neutral pH is obtained and equilibrated
with aqueous solution of KNO3 resulting in changing in composition and textural properties. As calculated
from N2 sorptiondesorption isotherms, surface specic area and cumulative pore volume for parent K-GP are
5.5 m2 g1 and 0.01 cm3 g1, respectively, whereas for pure K-GP surface specic area and cumulative pore
volume are 270 m2 g1 and 0.4 cm3 g1, respectively.
2013 Elsevier B.V. All rights reserved.

1. Introduction
Geopolymer is a name suggested for alkali-activated aluminosilicate binders capable of hardening at ambient conditions due to the formation of AlO
4 SiO4 tetrahedral framework (Davidovits, 1991). In the
framework formation, alkali or alkaline earth cations (Na+, K+, Ca2+
and in fewer cases Cs+) act as charge balancing species (O'Connor
et al., 2010; Provis et al., 2005a,b). Although the possibility of preparing
binding materials by the reaction of natural aluminosilicates with alkaline solutions goes back to the 1950s (Glukhovsky, 1957), some aspects
of the polycondensation mechanism are still the subjects of debate. The
structure of geopolymers synthesized at temperatures below 100 C is
dened as X-Ray amorphous; however, some authors mention the coexistence of nanometer-sized zeolitic crystals with long- and mediumrange disordered aluminosilicate gel (Bell et al., 2008; Palomo et al.,
1999; Provis et al., 2005a,b). During the last 30 years geopolymers
have been considered for a variety of applications, including an environmental friendly alternative to Portland cement (Duxson et al., 2007;
Hardjito et al., 2004; Palomo et al., 1999), as refractories (V. Barbosa
and MacKenzie, 2003; V.F.F. Barbosa and MacKenzie; Davidovits,
1991), ceramic precursors (Bell et al., 2009; He et al., 2010; Ikeda
et al., 2001), and matrices for immobilization of radioactive wastes
(Van Jaarsveld et al., 1998; Zosin et al., 1998), and as catalytic materials
(O'Connor et al., 2010; Sazama et al., 2011). Geopolymers can be also
viewed as amorphous analogs of zeolites, capable of cation exchange
and possessing catalytic properties when incorporated with Pt, Fe, Cu
and Co (Bortnovsky et al., 2008; Sazama et al., 2011). Although zeolitic
Tel.: +1 4802955791.
E-mail address: tskorina@mit.edu.
0169-1317/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.clay.2013.11.003

ion exchangers already hold a good portion of sorbent markets due to


their relatively low cost, cation exchange properties of geopolymers
have not been extensively studied: a comprehensive set of data has
been reported only for Na-based geopolymers (O'Connor et al., 2010).
Natural zeolites usually contain impurities, which do not take part in
ion exchange; therefore the theoretical ion exchange capacity (TEC) of
natural zeolites expresses the number of possible exchangeable cations.
For pure zeolites, TEC measured by elemental analysis is equal to
the negative charge of the aluminosilicate framework. Real exchange
capacity (REC) refers to the amount of actual exchangeable cations of
zeolite measured by ion-exchange methods (Inglezkis, 2005; Lehto
and Harjula, 1995). For some zeolites, TEC can be higher than REC for
certain pairs of cations e.g. for Cs+Na+ exchange in zeolite Y. Sodium
ions located in small cages of zeolite Y framework cannot be replaced
by the Cs+ ion which has a larger ionic radius, therefore this ion
exchange isotherm terminates at a point less than 100% (Sherry, 1966).
Cationic selectivity of an ion exchanger is generally determined by
measuring of an equilibrium uptake of ingoing cations from a solution.
Visual inspection of the ion exchange isotherms allows making conclusions as to the preference of an ion exchanger for the ingoing or
outgoing cation. The full description of the analysis of ion exchange
data can be found in the literature (Dyer, 1988; Sherry, 1966). Ionic
radius, hydration energy, and locations of cations (small cages or
super cages) are the main factors determining the preference of zeolite
for a cation over the other one. In zeolites, the mechanism of cation
attachment to the aluminosilicate framework also affects selectivity of
ion exchange. Hydrated and mobile cations act in a different way than
partially bare cations coordinated to oxygen atoms (Sherry, 1968). It appears that in the amorphous structure of geopolymers, negative charge
is not localized, and it is more or less uniformly distributed in the

206

T. Skorina / Applied Clay Science 87 (2014) 205211

framework (Duxson et al., 2007). Therefore, charge-balancing cations


can act as fully hydrated and mobile or as unhydrated (coordinated to
oxygen atoms by analogy with aluminosilicate glasses). For Na-based
geopolymer, it has been demonstrated by 23-Na NMR that the cations
are hydrated and they do not form part of the tetrahedral framework.
However, the 23-Na spectra shows that the Na environment is sensitive
to dehydration: heating of Na-based geopolymer up to 11001300 C
causes changes of typically hydrated sites to sites more typical of
unhydrated glasses (V. Barbosa and MacKenzie, 2003; V.F.F. Barbosa
and MacKenzie).
The main objective of this work has been to characterize K-based
geopolymer as an ion exchanger by determining the maximum cation
exchange level and ion exchange selectivity, taking into account structural and composition alteration associated with applied aqueous
treatment.

Isotherm points were obtained by contacting 0.5-g aliquots of dried


and ground pure K-GP with 50-cm3 aqueous solutions containing nitrates of the in-going cation (Cs+ or Na+) and the indigenous cation
(K+). The exchange solutions contained varying proportions of ingoing and indigenous ions in the range from 10 to 100 mmoldm3.
The system was stirred at 22 C for 24 h and subsequently equilibrated
for 48 h. Solid and liquid phases were separated, and the content of the
aqueous phase entrained in the solid was determined after equilibration. Both exchanging phases, the solid geopolymer and the exchanging
solution, were analyzed by PIXE and ICPMS (Thermo Scientic, USA).
The isotherms are expressed as Ag = f(As) at the constant total ion concentration of the solution (Nt, mmoldm3), where Ag and As are the
equivalent fractions of ion A in the geopolymer and solution. The subscript s refers to the solution and subscript g refers to geopolymer.
The equivalent fractions of the exchanging cation in solution and in
geopolymer are given as

2. Experimental
2.1. Synthesis of parent geopolymer
Parent potassium geopolymer samples (K-GP) were made by
dissolving potassium hydroxide (Sigma Aldrich, 85%) in deionized
water followed by addition of a proportional amount of fumed silica
(CAB-O-SIL EH-5, Cabot Corporation) along with metakaolin powder. Metakaolin powder was obtained by calcination of kaolinite
(Al2Si2O72H2O Alfa Aesar, N97% purity) at 800 C. This mixture was
stirred vigorously for about 10 min, and then cast into a plastic container.
The parent geopolymer was formulated with the following mole ratios:
Si/Al = 2.8, K/Al = 3, H2O/Al = 10. Polycondensation was performed
in a closed vessel in laboratory oven at 60 C for 24 h. After aging for
more than 1 week, the geopolymer samples were ground in an agate
mortar with subsequent separation of the fraction 1.21.6 mm.
2.2. Preparation of pure geopolymeric ion exchangers
Before the study of ion exchange properties, geopolymers, by analogy with zeolites, should have a good cation balance, i.e. the atomic ratio
of charge-balancing cations to aluminum is to be equal to 1 (Sherry,
1968). In this regard, it is extremely important to remove all soluble
species from parent geopolymer avoiding overwashing (replacement
of charge balancing ions by H3O+). Geopolymeric ion exchanger that
met these requirements (pure K-GP) was prepared as follows. About
1 g of the ground K-GP was placed into 50 cm3 centrifuge tube and repeatedly washed with deionized water until neutral pH is obtained
(usually 57 cycles). After the nal decantation, the tube containing
pure geopolymer was relled with 50 cm3 of 1 M KNO3. The sample
was treated under stirring (700 rpm) for 24 h at 22 C followed by
5 cycles of washing with 1 M solution and 5 cycles with 103 M solution of the same salt, then briey washed with deionized water and
dried at 60 C to the constant weight.
2.3. Preparation of maximally exchanged geopolymers
The samples of K-GP maximally exchanged with Na+ and Cs+
(labeled NaK-GP and CsK-GP, respectively) were prepared by exhaustively exchanging of 1 g aliquots of pure K-GP with 50 cm3 of 1 M
solution of the appropriate nitrate in a centrifuge tube tted with a
stirrer. The system was stirred for 24 h and then allowed to equilibrate
for 48 h at which point the geopolymer phase was separated by
centrifuging, briey washed with deionized water, and dried at 60 C
to the constant weight.
2.4. Construction of ion exchange isotherms
All isotherms were constructed at 22 C, at the constant ionic
strength (total exchange solution concentration of 100 mmoldm3).

As A=Nt
Ag mA =Mt ;
where [A] and mA are the equilibrium concentrations of ion A in solution
(mmoldm3) and in the geopolymer (mmolkg1) respectively, and
Mt is the theoretical ion exchange capacity of the geopolymer sample
determined by elemental analysis (mmolkg1).
The silica- and charge-balancing cations to aluminum mole ratios
were determined (Table 1). Elemental analysis of the solid phase was
performed by Proton Induced X-ray Emission (PIXE) using 1.7 MV
Tandetron Accelerator (General Ionex Corp). Proton beams were accelerated at 1.8 MeV. The proton beam crosses a 7 m kapton foil window
before entering the chamber containing the sample. Samples were
kept at 200 mTorr pressure for the measurement. The PIXE detector
(Si(Li)-type) signals were processed with the Windows GUPIX software
package.
2.5. Materials characterization
X-Ray powder diffraction (XRPD) patterns were recorded in
PANalytical X'Pert Pro Multipurpose Diffractometer with CuK radiation and X'Celerator position sensitive detector. The generator tension
and current were set at 45 kV and 40 mA, respectively. Samples were
scanned over 2 angles in the range 1565. The step size and scan
speed were set at 0.017 and 0.014/s, respectively. Data acquisition
was performed using the X'Pert Data Collector program, and the
XRD patterns were processed by the PANalytical X'Pert HighScore Plus
software (V3.0.5).
Nitrogen sorption isotherms were collected at a Micrometric
ASAP 2020 Surface area and Porosity Analyzer at 77 K. Samples were
degassed under vacuum at 150 C for ~8 h. For surface specic area calculation, the BraunauerEmmettTeller (BET) model has been applied
to the desorption branch. For pore size distribution, the Barrett
JoynerHalenda (BJH) model was applied to the desorption branch
with the Halsey thickness curve (Halsey, 1948) for non-uniform surfaces and Faas correction (Faas, 1981) to account for multilayer desorption in adsorbed layer thickness estimation. Total pore volume was
Table 1
Chemical composition of geopolymeric ion exchangers determined by PIXE.
Sample

Pure K-GP
CsK-GP
NaK-GP
a

Xmax, (%)a

Atomic ratios
Si/Al

K/Al

Na/Al

Cs/Al

2.11
2.20
2.23

1.02
0.39
0.23

0
0
0.77

0
0.61
0

The maximum exchanged level of the indigenous ion.

0
61
77

T. Skorina / Applied Clay Science 87 (2014) 205211

207

estimated from the total quantity of gas adsorbed at the data point
closest to P/P0 = 0.98 on the desorption branch.
Thermogravimetric analysis (TGA) study was carried out utilizing
a Mettler-Toledo TGA/DSC 1STARe system. Samples were analyzed
by heating from 25 to 1200 C at 10 C/min, holding at 1200 C for
30 min and nally cooling to room temperature in 30 min, all analyses
were carried out in air atmosphere.
For transmission electron microscopy (TEM) studies, samples were
prepared by grinding in ethanol to form dispersion. TEM grids were
dipped into the dispersion, taken out and dried in air. TEM images
were collected on a JEOL 2010 at the accelerating voltage of 200 kV.
CLSM images of methylene blue-impregnated pure K-GP (210 m
fraction) were collected on a Zeiss LSM-510 confocal laser-scanning
microscope. A 100/1.4 oil immersion objective with numerical aperture of 1.4 was used. The 543 nm line of a HeNe laser was used for
the excitation of the dye molecules embedded in geopolymer matrix.
3. Results and discussions
The results of the elemental analysis of pure K-GP and the maximally
exchanged forms (NaK-GP and CsK-GP) are shown in Table 1. The
molar ratio of charge-balancing cations to aluminum was 1 0.02 in
each case, which indicates that degree of H3O+ exchange was negligible.
However, the silica to alumina molar ratio (SAR) increases during
the treatment due to hydrolytic degradation of Si\O\Al framework.
Generally, ultimate properties of a geopolymer, including hydrolytic
stability, strongly depend on the formulation of the precursors and
curing conditions i.e. duration and temperature. Despite curing of
our geopolymer was performed at the conditions widely accepted for
this synthesis (60 C, 24 h), the instability of SAR might reect still incomplete reaction between the metakaolin and alkalisilica precursors.
In terms of quantitative analysis of ion-exchange data, the instability
of SAR implies that the maximum exchanged level should be expressed
as the percentage of charge-balancing ions replaced by the in-going ions
(Xmax, %), which allows the comparison of the samples with different
SAR. The maximum percentage of the indigenous cations (K+) replaced
by Na+ is moderately higher than that replaced by Cs+. This indicates
the geopolymer preference for cation with smaller ionic radius, but
complete removal of indigenous cations was not achieved either.

Fig. 2. TGA curve for pure KGP after drying at 60 C, air atmosphere, 10 C/min.

Long- and medium-range disordered geopolymers are metastable in


respect to well-ordered zeolites with identical chemical composition.
Formation of zeolites (sodalite, cancrinite, and Na-chabazite,) as
the result of prolonged curing (from several days to several months)
of sodium-based geopolymeric mixture and/or mild hydrothermal
treatment is well documented (Criado et al., 2007; De Silva and
Sagoe-Crenstil, 2008; Desbats-Le Chequer and Frizon, 2011). However,
to the author's knowledge, there is no direct evidence of potassium
geopolymerzeolite transformation mediated by ion-exchange at ambient conditions (O'Connor et al., 2010). Taking into account that
secondary phase precipitation from the oversaturated batch solution
might be possible in the vicinity of any aluminosilicate surface
(Chorover et al., 2003; Hellmann et al., 2012), structural evolution of

Fig. 1. (a): Powder XRD patterns of parent geopolymer (K-GP), (b): pure K-GP after heating at 1100 C.

208

T. Skorina / Applied Clay Science 87 (2014) 205211

geopolymer accompanied the repeated treatment with aqueous solutions of nitrates was carefully monitored.
The XRD pattern of parent geopolymer (Fig 1a) and its exchanged
forms (not shown here due to their identity) indicate typical X-Rayamorphous structure. The diffuse peak centered at 2 Theta 30 has
been considered in the literature as the distinguishing feature of
geopolymers; this hump can be attributed to nano-sized crystalline
aluminosilicates e.g. zeolites (Bell et al., 2008; Provis et al., 2005a,b).
The absence of amorphous peak centered at 2 Theta = 22 C indicates
a negligible presence of metakaolin (Davidovits, 1991; Rowles and
O'Connor, 2003). Sharp diffraction peaks indicated with A demonstrate the presence of an anatase (PDF#: 04-011-0664), which is an impurity commonly found in Kaolinite. A trace amount of crystalline
cesium aluminum nitrate (PDF#: 04-011-0664) was found in a number
of studied samples, but, it appears that this formation is not stable to
the batch purication. Fig. 2 shows TGA and DSC curves obtained in
the temperature range 251200 C for pure K-GP preliminary dried
at 60 C. The weight loss observed below 250 C (~11 wt.%) is related

to the evaporation of physisorbed (residual) and capillary water, this


period also includes the initial elimination of hydroxyl groups associated with the surface. The dehydration in the temperature range
250600 C reveals further elimination of hydroxyl groups (~ 2 wt.%)
associated with less available surface of aluminosilicate gel. The weight
loss observed from 6001200 C is negligible. A small exothermic peak
registered by DSC at 1100 C is related to leucite (PDF#: 01-085-1421)
crystallization conrmed by XRPD of pure K-GP recorded after 24 hour
heating at 1100 C (Fig. 1b). The relatively low crystallization temperature of leucite is believed to be the result of the presence of nuclei in
the geopolymers, which form during initial polycondensation (Bell
et al., 2009).
The textural properties of parent K-GP, pure K-GP and maximally
exchanged NaK-GP and CsK-GP forms were determined from the
analysis of N2 sorptiondesorption isotherms using BET and BJH models
(Table 2). It should be emphasized that the detectable range of pore size
is 2 nm500 nm for this technique, which is why the macro porosity of
geopolymer is not discussed in this section. It has been observed that
BET surface area and total pore volume for pure K-GP are substantially
higher than those for the parent geopolymer and the maximally
exchanged forms.
It has been repeatably observed that the textural properties of
geopolymer are signicantly altered after equilibration with solution
of parent cation nitrate (KNO3) followed by washing with deionized
water. It should be noted that samples washed with deionized water
until neutral pH is obtained exhibit similar textural properties (not
shown here). However, such substantial values of surface area and
total pore volume are not retained after subsequent equilibration of
the sample with foreign ions; Na- and Cs-forms of pure K-GP exhibit
even lower surface specic area and pore volume than those for parent

Fig. 3. (a): N2-sorbtion isotherms for pure K GP (), and (b): parent K-GP (), CsK-GP
(), NaK-GP ().

Fig. 4. (a): BJH pore size distribution for pure K-GP (), and (b): parent K-GP (), CsK-GP
(), NaK-GP ().

Table 2
The textural properties of geopolymeric ion exchangers.
Sample

BET surface
area (m2g1)

T-Plot external
surface area a
(m2g1)

Total pore
volume
(cm3 g1)

Average pore
size b (nm)

Initial K-GP
Pure K-GP
CsK-GP
NaK-GP

5.5
268
1.7
2.8

4.90
261
1.00
2.32

0.01
0.40
0.01
0.01

9.0
5.7
20.7
19.4

a
b

Surface area of pores larger than 2 nm.


4(Total pore volume)/(BET surface area).

T. Skorina / Applied Clay Science 87 (2014) 205211

K-GP. Although the values of SSA and the total pore volume for pure
K-GP are an order of magnitude higher than those for parent K-GP,
the character of nitrogen sorption isotherms of pure K-GP (shown
Fig. 3a) and parent K-GP (shown in Fig. 3b) is similar. Both demonstrate
a distinguishable H2 type sorptiondesorption hysteresis indicating
the presence of mesopores and possible pore interconnectivity. The
pure K-GP demonstrates a quite narrow BJH pore size distribution
(Fig. 4a) characterized by the predominance of small mesopores of
about 5 nm width. The maximally exchanged CsK-GP and NaK-GP
exhibit typical Type II isotherms with negligible sorptiondesorption
hysteresis loop (Fig. 3b). This type of isotherms is frequently encountered for sorption on the structures that do not contain interconnecting
channels and mesopores. The predominance of macropores in the structure of the maximally exchanged geopolymer is also conrmed by the
BJH pore size distribution plotted in Fig. 4(b).
Penetration prole of dye under room temperature into particles of
pure K-GP has been studied by CLSM. The examination has revealed
that methylene blue molecules unable to penetrate deep along the
porous matrix and remain adsorbed on the external surface only.
The depth of the penetration prole varies from hundreds of nanometers to several microns (Fig. 5a,b). TEM micrographs obtained for
ground pure K-GP exhibit microstructures typical for conventional
geopolymers. As indicated in Fig. 5 (c,d) the size of primary particles
and intra-structural pores in pure K-GP ranges from 10 to 20 nm and
2050 nm, respectively. The microscopic observations are consistent
with BET data that register quite high accessible surface area and the
presence of pores 1020 nm, but do not show interconnectivity.

209

The isotherms for the alkali metal ion exchange with the potassiumbased geopolymer at 22 C are shown in Fig 6. Cesium exchange of pure
K-GP (in-going cation has larger size than indigenous ion) shows a
denite termination at 61% of the theoretical capacity. Moreover, for
pure K-GP, the complete replacement of indigenous ions has not been
achieved also for Na+K+ exchange, when in-going cation has smaller
size than indigenous ion. There is a general agreement that potassium,
rubidium and cesium ions are water structure breakers, and that
lithium ion is a water structure maker. Whether a second hydration
shell around the sodium ion is present or not had been the subject of
debate for a long time. However, the double difference infrared spectroscopy study reported recently (Mhler and Persson, 2012) indicates
a second hydration sphere for Li+ but not for Na+. The ionic radii proposed by these authors for ve- and six-coordinate sodium are 1.02
and 1.07 , respectively. The ionic radii for six- and seven-coordinate
K+ are 1.38 and 1.46 , respectively, and eight-coordinate Rb+ and
Cs+, 1.64 and 1.73 , respectively. Thus, sodium ion has a smaller radius
than potassium ion both being bare and hydrated, and hence, the observation that Na+K+ exchange in pure K-GP is incomplete clearly shows
partial accessibility of charge-balancing ions, previously observed in
(Bortnovsky et al., 2008).
In contrast, it is well documented that for zeolite Y, complete
replacement of the indigenous alkali metal ion can be achieved by
exhausted ion exchange with Li+, Na+, and K+. Cesium and rubidium
ions cannot penetrate through 6-ring connecting windows and, therefore, only capable of partial exchange (~70%), regardless of the hydration state of the unit cell (Barrer et al., 1966; Sherry, 1968). This

Fig. 5. (a): Confocal laser scanning microscopy cross-sectional images of pure K-GP impregnated with dye (false color), (b): uorescence intensity map produced by sorbed dye;
(cd): transmission electron microscopy images of pure K-GP.

210

T. Skorina / Applied Clay Science 87 (2014) 205211

This re-ordering of the selectivity series demonstrates that at higher


loadings bare or partially bare ions coordinated to framework oxygen
atoms participate in exchange process (Sherry and Howard, 2003).
In order to see selectivity trend in the isotherms associated with incomplete exchange, a normalization procedure is required, in which
the experimental Ag values are multiplied by a normalization factor
f = 1/Xmax, where Xmax represents the maximum percentage of ions
exchanged by in-going ion A (Barrer and Klinowski, 1972; Keane,
1996). As can be seen from our normalized Cs/K-GP isotherm
shown in the insert in Fig. 6(a), pure K-GP exhibits a slight preference
for K+ over Cs+, which becomes more distinguishable with increase
in Cs+ loading. The normalized Cs+Na+ isotherm for NaK-GP
shown in the insert part of Fig. 6(b) winds around the selectivity line indicating a poorly distinguishable preference of geopolymer for Na+ over
Cs+ at low Na+ loading and a converse preference at high loading.
4. Conclusion

Fig. 6. Alkali metal ion exchange isotherms associated with incomplete exchange obtained
at 22 C and 0.1 total normality for the exchange of (a) pure K-GP, and (b) NaK-GP;
() Cs+ and () Na+; inserts show the isotherms normalized by maximum exchanged
levels highlighting the selectivity trends.

difference in ion exchange behavior of zeolites and geopolymers can


be understood if the interconnectivity of pores is considered. In the
exhaustive exchange, when a considerable concentration difference between phases is applied, all mobile cations diffuse out from the matrix.
This process might be prevented by either the pores being too small
(encapsulation), or the pores being present, but not connected to the
outside. The latter is inherent in the non-regular nature of geopolymers
and extrinsic to the well-ordered zeolites.
As mentioned earlier, ion selectivity is typically affected by the
bonding mechanism of charge-balancing cations. In geopolymer this
mechanism is still not fully understood, but speculated to be the
following: a partially bear cation is either bonded into the matrix
via `Al\O\ sites and/or being hydrated, present in the framework
cavities to balance the delocalized negative electrical charge. For
zeolites, the existence of hydration shell for charge-balancing cations
located in the large cages has been indirectly assessed using the
experimental selectivity series (Sherry, 1968). For instance, at low
loading (b 50%) of the incoming ion, the selectivity series for zeolite
X and Y is Cs N Rb N K N Na, indicating that only mobile, hydrated
ions in the large cages are replaced. In contrast, at higher loadings
5060%, the selectivity of zeolite X decreases with increasing ionic
radius of ingoing ion (except Li+ that is water structure maker).

In this work, the maximum exchange level of potassium-based


geopolymer has been determined for the exchange with Cs+ and Na+
(61 and 77% respectively). The incomplete exchange observed for potassium based geopolymer under the exhaustive conditions indicates
partial accessibility of charge-balancing ions both for in-going cations
of smaller and larger radii than those for the ingenious cations.
A slight preference of the studied geopolymer for K+ over Cs+ under
noncompetitive exchange has been shown by the normalized Cs/K-GP
isotherm. The BET surface specic area and pore volume of the studied
geopolymer have been found to drastically increase during the repeated
equilibration with solution of KNO3 or repeated washing with deionized
water under room temperature. BET surface specic area observed for
parent geopolymer = 5.5, cumulative pore volume = 0.01 cm3 g1,
whereas surface specic area of pure K-GP = 270 m2 g1, and pore
volume = 0.4 cm3 g1. It has been demonstrated herein the possibility
of preparation of highly porous sorbents capable of almost non-selective
cation exchange based on conventional geopolymers. It should also be
emphasized that the synthesis and proposed treatment of geopolymers
can be considered as environmental friendly processes, which, unlike
zeolite syntheses, do not require high temperature and elevated pressure.
Furthermore, the geopolymeric sorbents can be synthesized utilizing
waste products, such as slags and y ash.
Acknowledgments
Most of the experimental work has been done at the Department of
Chemistry and Biochemistry, Arizona State University, US. The author
is grateful to Dr. Don Seo for the scientic guidance and providing lab
facilities and equipment.
References
Barbosa, V., MacKenzie, K., 2003a. Synthesis and thermal behavior of potassium sialate
geopolymers. Mater. Lett. 57 (910), 14771482.
Barbosa, V.F.F., MacKenzie, K.J.D., 2003b. Thermal behavior of inorganic geopolymers and
composites derived from sodium polysialate. Mater. Res. Bull. 38, 319331.
Barrer, R.M., Klinowski, J., 1972. Ion exchange involving several groups of homogeneous
sites. J. Chem. Soc. Faraday Trans. 1 (68), 7387.
Barrer, R., Rees, L., Shamsuzzoha, J., 1966. Thermochemistry and thermodynamics of ion
exchange in a near-faujasite. J. Inorg. Nucl. Chem. 28, 629643.
Bell, J., Sarin, P., Driemeyer, P., Ryan, P., Haggerty, R., Chupas, P., Kriven, W., 2008. X-Ray
pair distribution function analysis of a metakaolin-based, KAlSi2O6 5.5H2O inorganic
polymer (geopolymer). J. Mater. Chem. 18, 59745981.
Bell, J., Driemmeyer, P., Kriven, W., 2009. Formation of ceramics from metakaolin-based
geopolymers: Part II: K-based geopolymer. J. Am. Ceram. Soc. 92 (3), 607615.
Bortnovsky, O., Ddeek, J., Tvarkov, Z., Sobalk, Z., 2008. Metal ions as probes for characterization of geopolymer materials. 91 (9), 30523057.
Chorover, J., Choi, S., Amistadi, M., Karthikeyan, K., Crosson, G., Mueller, K., 2003. Linking
cesium and strontium uptake to kaolinite weathering in simulated tank waste
leachate. Environ. Sci. Technol. 37 (10), 22002208.
Criado, M., Fernndez-Jimnez, A., De la Torre, A., Aranda, M., Palomo, A., 2007. An XRD
study of the effect of the SiO2/Na2O ratio on the alkali activation of y ash. Cem.
Concr. Res. 37 (5), 671679.

T. Skorina / Applied Clay Science 87 (2014) 205211


Davidovits, J., 1991. Geopolymers inorganic polymeric new materials. J. Therm. Anal.
Calorim. 37, 16331656.
De Silva, P., Sagoe-Crenstil, K., 2008. Medium-term phase stability of Na2OAl2O3SiO2H2O
geopolymer systems. Cem. Concr. Res. 38 (6), 870876.
Desbats-Le Chequer, C., Frizon, F., 2011. Impact of sulfate and nitrate incorporation
on potassium-and sodium-based geopolymers: geopolymerization and materials
properties. J. Mater. Sci. 46, 56575664.
Duxson, P., Fernndez-Jimnez, A., Provis, J., Lukey, G., Palomo, A., Van Deventer, J., 2007.
Geopolymer technology: the current state of the art. J. Mater. Sci. 42, 29172933.
Dyer, A., 1988. An Introduction to Zeolite Molecular Sieves. John Wiley & Sons Inc.,
New York.
Faas, G., 1981. Correlation of Gas Adsorption, Mercury Intrusion, and Electron Microscopy
Pore Property Data for Porous Glasses. (Thesis) Georgia Institute of Technology.
Glukhovsky, V., 1957. A new building material. Bulletin of Technikal Information,
Glavkievstroy, N2.
Halsey, G., 1948. Physical adsorption on nonuniform surfaces. J. Chem. Phys. 16, 931937.
Hardjito, D., Wallah, S., Sumajouw, D., Rangan, B., 2004. On the development of y ashbased geopolymer concrete. ACI Mater. J. 101 (6), 467472.
He, P., Jia, D., Lin, T., Wang, M., Zhou, Y., 2010. Effects of high-temperature heat treatment
on the mechanical properties of unidirectional carbon ber reinforced geopolymer
composites. Ceram. Int. 36 (4), 14471453.
Hellmann, R., Wirth, R., Daval, D., Barnes, J., Penisson, J., Tisserand, D., Hervig, L., 2012. Unifying natural and laboratory chemical weathering with interfacial dissolution
reprecipitation: a study based on the nanometer-scale chemistry of uidsilicate
interfaces. Chem. Geol. 294, 203216.
Ikeda, K., Onikura, K., Nakamura, Y., Vedanand, S., 2001. Optical spectra of nickel-bearing
silicate gels prepared by the geopolymer technique, with special reference to the
low-temperature formation of liebenbergite (Ni2SiO4). J. Am. Ceram. Soc. 84,
17171720.
Inglezkis, V., 2005. The concept of capacity in zeolite ion-exchange systems. J. Colloid
Interface Sci. 281, 6879.

211

Keane, M., 1996. The role of the alkali metal co-cation in the ion exchange of Y zeolites IV.
Cerium ion exchange equilibria. Microporous Mater. 7 (1), 5159.
Lehto, J., Harjula, R., 1995. Experimentation in ion exchange studies the problem of
getting reliable and comparable results. React. Funct. Polym. 27 (2), 121146.
Mhler, J., Persson, I., 2012. A study of the hydration of the alkali metal ions in aqueous
solution. Inorg. Chem. 51 (1), 425438.
O'Connor, S.J., MacKenzie, K.J.D., Smith, M., Hanna, J., 2010. Ion exchange in the chargebalancing sites of aluminosilicate inorganic polymers. J. Mater. Chem. 20, 1023410240.
Palomo, Grutzeck, M., Blanco, M., 1999. Alkali-activated y ashes. A cement for future.
Cem. Concr. Res. 29, 13231329.
Provis, J., Lukey, G., van Deventer, Jannie S.J., 2005a. Do geopolymers actually contain
nanocrystalline zeolites? A reexamination of existing results. Chem. Mater. 17,
30753085.
Provis, J., Duxson, P., Lukey, G., van Deventer, Jannie S.J., 2005b. Statistical thermodynamic
model for Si/Al ordering in amorphous aluminosilicates. 17 (11), 29762986.
Rowles, M., O'Connor, B., 2003. Chemical optimisation of the compressive strength of aluminosilicate geopolymers synthesised by sodium silicate activation of metakaolinite.
J. Mater. Chem. 13, 11611165.
Sazama, P., Bortnovsky, O., Ddeek, J., Tvarkov, Z., Sobalk, Z., 2011. Geopolymer
based catalysts new group of catalytic materials. Catal. Today 164, 9299.
Sherry, H., 1966. The ion-exchange properties of zeolites. I. Univalent ion exchange in
synthetic faujasite. J. Phys. Chem. 70 (4), 11581168.
Sherry, H., 1968. Ion-exchange properties of zeolites. IV. Alkaline earth ion exchange in
the synthetic zeolites Linde X and Y. J. Phys. Chem. 72 (12), 40864094.
Sherry, H., Howard, S., 2003. Ion exchange. Handbook of Zeolite Science and Technology
10071061 Marcel Dekker, New York.
Van Jaarsveld, J.G.S., Van Deventer, J.S.J., Lorenzen, L., 1998. Factors affecting the immobilization of metals in geopolymerized y ash. Metall. Mater. Trans. B 1, 283291.
Zosin, A., Priimak, T., Avsaragov, Kh., 1998. Geopolymer materials based on magnesia
iron slags for normalization and storage of radioactive wastes. Atom. lnergiya 85
(1), 7882.

You might also like