You are on page 1of 7

American Mineralogist, Volume 86, pages 431–437, 2001

The effects of time, temperature, and concentration on Sr2+ exchange in clinoptilolite in


aqueous solutions

JENNIFER L. PALMER AND MICKEY E. GUNTER*

Department of Geological Sciences, University of Idaho, Moscow, Idaho 83844-3022, U.S.A.

ABSTRACT
Four grams of a clinoptilolite-rich rock, crushed to –180 mesh, were exchanged in deionized
water (DI), and in 0.1, 0.01, and 0.001 M solutions of SrCl2·6H2O at 5, 21, 50, and 90 °C for 0.5, 5,
24, and 240 h yielding 64 sample solutions, which were vacuum-filtered to remove solids. The
solutions were analyzed by ICP-AES for the outgoing cations (i.e., Na+, K+, Ca2+, and Mg2+). The
higher molarity solutions were expected to contain greater concentrations of the outgoing cations
than the DI or lower molarity concentration solutions. This trend was exhibited by the monovalent
and divalent cations in samples of all concentrations. Outgoing monovalent cations exchanged to
greater extents with increasing time and temperature. Outgoing divalent cations of 0.01 and 0.001
M concentrations exchanged at low temperatures contained greater concentrations of outgoing di-
valent cations than solutions exchanged at high temperatures. The concentrations of the monova-
lent cations were slightly greater in the 0.001 M sample than in DI, whereas divalent cations had a
similar concentration between the 0.001 M and DI treatments. Overall exchange of cations out of
the clinoptilolite was favored at higher temperatures.

INTRODUCTION
Zanetti and Gunter 1997). This study demonstrates the effects
Due to its cation exchange capabilities and availability in of time, temperature, and Sr concentration on the release of
mineable quantities, clinoptilolite has been useful in industry cations from clinoptilolite upon Sr exchange.
for removal of metals from contaminated soils and solutions The main goal of this experiment was the same as that of
(Leppert 1990; Las 1989; Mumpton 1978). Because of the se- Woods and Gunter (2001): to examine how different cations
lectivity of clinoptilolite for Sr and Cs over other cations (Ames (Na, K, Ca, and Mg) exchange out of clinoptilolite and into a
1960, 1961), the mineral has been used to remove radioactive liquid of homoionic composition, in this case solutions of SrCl,
Sr and Cs from low-level radioactive waste (Las 1989; Baxter and not to perform a thermodynamic study like that done by
and Berghauser 1986; Mumpton 1978). The Site Ion Exchange Barkatt and Van Konynenbugh (1993), for example. To our
Effluent Plant of British Nuclear Fuels in Sellafield, UK cur- knowledge, these types of exchange experiments have not been
rently employs clinoptilolite for the removal of 90Sr and 134,137Cs previously performed. Researchers routinely pretreat
from low-level radioactive cooling-pond waters before dis- clinoptilolite in an NaCl solution and then examine how the
charge into the Irish Sea (Las 1989; Baxter and Berghauser homoionic form of the solid behaves in aqueous solutions of
1986). Although the pond water contains more Cs than Sr (by either single or multiple cations. These types of experiments
an order of magnitude), Sr breakthrough occurs before Cs break- are very important in understanding the thermodynamics of
through (i.e., Sr stops exchanging with clinoptilolite before Cs) exchange and various industrial applications. However, zeo-
and is the limiting factor in the life of the exchange beds. An lites in the natural environment will not be homoionic but will
understanding of the factors governing Sr and Cs exchange with contain several cations. So the obvious question is whether these
clinoptilolite is necessary for developing ways to prolong the cations all exchange out of the solid and into the liquid in the
life of the beds. same manner or different manners. Thus, our main goal in this
Many studies have been performed on the exchange of study was to determine the behavior of these outgoing cations
clinoptilolite for Sr and Cs by examining the Sr and Cs content (i.e., effect of temperature, release rate, and time) for Na, K,
of the mineral (Las 1989; Ames 1960, 1961). However, the Ca, and Mg in a clinoptilolite to better understand cation (and
behavior of the outgoing cations is poorly understood. The ef- water) exchange into and out of natural samples.
fects of time and temperature on Cs exchange in clinoptilolite
have been determined in other studies (Woods and Gunter 2001; CATION EXCHANGE IN CLINOPTILOLITE
Clinoptilolite, (Na,K)6(Al6Si30O72)·20H2O, initially was ac-
cepted as a lower symmetry, monoclinic form of ptilolite
* E-mail: mgunter@uidaho.edu (Schaller 1932). Subsequently, clinoptilolite was assigned to
0003-004X/01/0004–431$05.00 431
432 PALMER AND GUNTER: ION-EXCHANGE IN CLINOPTILOLITE-RICH ROCKS, PART II

the heulandite group and a debate ensued about its indepen-


dence as a mineral species separate from heulandite,
(Na,K)Ca4(Al9Si27O72)·24H2O (Coombs et al. 1997; Boles 1972;
Mason and Sand 1960; Mumpton 1960; Hey and Bannister
1934). The two minerals most recently have been differenti-
ated by their Si/Al ratios, with clinoptilolite having Si/Al > 4,
and heulandite Si/Al < 4 (Coombs et al. 1997). In general, heu-
landite occurs as a secondary mineral found within vugs of
basalts, whereas clinoptilolite can occur in widespread forma-
tions of nearly pure microcrystalline form as an alteration prod-
uct of volcanic glass (Hay 1981; Surdam 1981). Despite the
chemical and structural similarities between heulandite and
clinoptilolite, the difference in their occurrences makes
clinoptilolite more available in mineable quantities, whereas
vugs in basalts provide large single crystals of heulandite for
structural studies. Therefore, the economic importance of
clinoptilolite has stimulated research on the chemistry, struc-
ture, and cation exchange properties of both heulandite and
clinoptilolite.
The structures of heulandite and clinoptilolite exhibit dif-
ferences in thermal stability. Upon heating between approxi-
FIGURE 1. The structure of natural heulandite (a) from Gunter et
mately 200–450 °C, heulandite experiences two phases of
al. (1994) as seen down [001]. The A and B channels are composed of
contraction before becoming amorphous, whereas clinoptilolite 10- and 8-membered tetrahedral rings, respectively, and are parallel to
can retain its structural integrity up to about 700 °C, a charac- c and the C channel is composed of 8-membered rings and is parallel
teristic that has been attributed to its higher silica content (Boles to a. Sodium, K, and Ca occupy sites in the A, B, and C channels with
1972; Mumpton 1960). However, in general, the two minerals their associated water molecules. The small amount of Mg that occurs
are structurally the same. Heulandite-group minerals conform in clinoptilolite would most likely occur in the C channel near the K
to the monoclinic space group C2/m with a ≈ 17, b ≈ 17, c ≈ 7 site. (b) A fully exchanged Sr sample of heulandite from Stolz and
Å, and β ≈ 116°. Heulandite and clinoptilolite are composed of Armbruster (2000). Sr occupies sites in the A and B channels with its
an aluminosilicate framework that exhibits a net negative associated water molecules.
charge, and contains a three-dimensional network of open chan-
nels into and out of which cations exchange to compensate for
and charge of the cation. If the water molecules move into and
the negative charge by bonding to framework oxygen atoms.
out of the framework channels with their associated cations,
The channels are defined by 8- and 10-membered tetrahedral
namely the hydration spheres move about independently,
rings that are stacked parallel to a (C channel) and c (A and B
changes in the structure of heulandites can be attributed to the
channels). Figure 1a illustrates the structure of a natural heu-
hydration spheres rather than to individual cations. This would
landite in which cations occupy sites in the A, B, and C chan-
suggest that cation exchange in clinoptilolite can be character-
nels (Gunter et al. 1994). Figure 1b illustrates a Sr-exchanged
ized by the behavior of the hydration spheres as they move
heulandite in which Sr2+ occupies sites in the A and B channels
into and out of the channels. Additionally, the hydration sphere
(Stolz and Armbruster 1997).
radius may change with temperature due to weak electrostatic
The crystal structure of heulandite minerals changes slightly
attraction of the cation and water molecules.
upon exchange of cations. For instance, upon incorporation of
Monovalent cations holding water molecules weakly can
Cs+ into its framework channels, heulandite experiences a
shed those molecules easily in order to fit into a site. Higher-
change in space group from C2/m to C1 (Yang and Armbruster
charged, smaller-sized cations, however, are bonded more
1996). Heulandite has also been known to experience de-
strongly to the water molecules and hence will shed them less
alumination upon exchange in acidic solutions (Stolz and
readily. In general, there is an inverse relationship between
Armbruster 1997; Wüst et al. 1999; Las 1989). Incorporation
cation radius and hydrated cation radius. Divalent cations are
of H+ from acidic solutions was predicted by Hey (1930) and
smaller than monovalent cations but usually have a larger hy-
deduced based on other structural (Stolz and Armbruster 1997;
drated radius than monovalent cations (Nightingale 1959).
Wüst et al. 1999) and optical (Palmer and Gunter 2000) work.
Changes in temperature and the charge of the exchangeable
The changes in the structure of heulandites are attributed to the
cation could therefore make clinoptilolite more or less selec-
cations that the mineral incorporates into its framework chan-
tive for that cation. For instance, Ames (1960) observed that
nels. However, the cations are known to be coordinated by water
the affinity of clinoptilolite for Cs decreased as temperature
molecules, forming clusters (Stolz and Armbruster 1997; Wüst
increased when Na was a competing cation, suggesting that
et al. 1999; Yang and Armbruster 1998; Yang and Armbruster
the hydrated radius of Na+ at higher temperatures approaches
1996; Armbruster and Gunter 1991; Semmens and Seyfarth
that of Cs+, thereby making Na more competitive for sites in
1978; Boles 1972; Ames 1961). The characteristics of these
the mineral with higher temperatures.
clusters, referred to as hydration spheres, depend on the size
PALMER AND GUNTER: ION-EXCHANGE IN CLINOPTILOLITE-RICH ROCKS, PART II 433

EXPERIMENTAL METHODS were chosen to demonstrate the initial rapid exchange followed
A clinoptilolite-rich tuff from the Mud Hills formation near by a decrease in the exchange rates. In fact, near steady-state
Barstow, California was used in this study. The rock consisted was reached after 240 h, so the 720 h data is not shown. After
of (about) 95% clinoptilolite with minor amounts of mica, gyp- the allotted storage time, the mixtures were vacuum filtered
sum, dolomite, albite, potassium feldspar, quartz, barite, mag- with a fritted glass filter fitted with a silica-glass microfiber
netite, montmorillonite, and illite (Sheppard and Gude 1969); filter to remove the solid from solution. The resulting solu-
however, we found no other crystalline phases in our sample tions were further filtered with 0.1 µm nylon membrane sy-
by powder X-ray diffraction (XRD). The chemical formula for ringe filters to remove any remaining particulate matter. The
the clinoptilolite, Na3.13K1.04Ca0.68Mg0.47(Al6.09Si29.60O72)·28H2O, final samples were preserved with 2 mL of trace metal-grade
was determined in a previous study using a Rigaku 3370 X-ray nitric acid. The resultant sample solutions were analyzed with
fluorescence spectrometer assuming that the rock was pure a Perkin-Elmer Optima 3000 XL Inductively Coupled Plasma
clinoptilolite (Zanetti and Gunter 1997). Atomic Emission Spectrometer (ICP-AES) at the University
The clinoptilolite was crushed and sieved to –180 mesh (<84 of Idaho for Na+, K+, Ca2+, and Mg2+ to determine the concen-
µm). 100 mL solutions of 0.1, 0.01, and 0.001 M SrCl2·6H2O trations of cations displaced from the mineral during exchange.
were mixed with 4 g of clinoptilolite, placed in 125 mL sealed The instrumental detection limits for these cations are 0.2, 0.3,
Nalgene containers, and treated at 5, 21, 50, and 90 °C for 0.5, 0.01, and 0.01 ppm respectively. An in-house concentration veri-
5, 24, 240, 720 h. The 5 °C sample was stored in a lab refrig- fication standard (CCV) was analyzed approximately every
erator, the 21 °C sample at room temperature, and the 50 and tenth sample. A boron internal standard was used to monitor
90 °C samples were stored in a lab oven. The time intervals instrumental drift. We did not measure Sr concentrations be-

TABLE 1. Results of ICP-AES analyses for concentrations (ppm) 24 h samples 5°C


of cations in solution for differing treatments 0.1 424(42) 6.8(7) 10(1) 41(4) 943
0.01 151(24) 2.4(2) 5.7(6) 12(1) 338
M Na K Mg Ca Sr
0.001 68(7) 0.88(9) 1.1(1) 2.7(3) 140
0.5 h samples 5°C DI 47(5) 0.34(4) 0.44(4) 0.88(9)
0.1 239(24) 4.0(8) 8.6(9) 21(2) 538
0.01 97(13) 1.3(1) 3.9(4) 7.7(8) 217 24 h samples 21°C
0.001 54(5) 0.8(2) 0.72(7) 2.1(2) 110 0.1 715(72) 34(3) 12(1) 70(7) 1597
DI 27(3) 0.73(7) 0.29(3) 0.61(6) 0.01 236(42) 11(1) 5.5(5) 14(1) 511
0.001 70(13) 1.7(2) 0.86(9) 2.3(2) 144
0.5 h samples 21°C DI 50(5) 0.52(7) 0.28(3) 0.44(4)
0.1 270(27) 5.1(5) 10(1) 26(3) 611
0.01 120(22) 1.9(2) 4.1(4) 7.9(8) 263 24 h samples 50°C
0.001 55(6) 1.0(1) 0.70(7) 2.3(2) 114 0.1 831(138) 124(12) 22(2) 304(30) 2465
DI 31(3) 0.92(9) 0.32(3) 0.67(7) 0.01 447(45) 24(2) 5.2(5) 18(2) 936
0.001 89(15) 2.5(2) 0.20(2) 0.40(4) 175
0.5 h samples 50°C DI 65(7) 1.6(2) 0.35(4) 0.47(5)
0.1 287(29) 8.0(8) 9.0(9) 24(2) 642
0.01 98(16) 2.3(2) 3.2(3) 6.5(6) 215 24 h samples 90°C
0.001 55(6) 1.0(1) 0.60(6) 2.1(2) 113 0.1 932(170) 259(26) 109(11) 455(46) 3454
DI 26(4) 0.85(8) 0.28(3) 0.59(6) 0.01 555(56) 41(4) 4.9(5) 14(1) 1152
0.001 115(21) 4.0(4) 0.17(2) 0.50(5) 225
0.5 h samples 90°C DI 80(8) 4.2(4) 0.87(9) 1.4(1)
0.1 450(45) 23(2) 10(1) 32(3) 990
0.01 121(20) 4.7(5) 3.0(3) 6.6(7) 261 240 h samples 5°C
0.001 59(6) 1.9(2) 0.63(6) 2.2(2) 121 0.1 657(66) 18(2) 12(1) 111(11) 1557
DI 28(3) 1.2(1) 0.36(4) 0.60(6) 0.01 404(40) 6(1) 8.3(8) 24(2) 859
0.001 93(15) 1.1(1) 0.85(8) 1.5(2) 185
5 h samples 5°C DI 47(5) 0.93(9) 0.37(4) 0.52(5)
0.1 326(33) 5.3(5) 10(1) 35(3) 739
0.01 108(14) 1.7(2) 4.4(4) 8.7(9) 242 240 h samples 21°C
0.001 58(6) 0.77(8) 0.89(9) 2.6(3) 120 0.1 749(75) 70(7) 16(2) 153(15) 2000
DI 27(3) 0.80(8) 0.40(4) 0.77(8) 0.01 463(46) 13(1) 8.3(8) 25(3) 982
0.001 102(17) 1.6(2) 0.46(5) 0.85(9) 200
5 h samples 21°C DI 65(7) 2.0(2) 0.55(6) 0.63(6)
0.1 527(197) 19(2) 11(1) 42(4) 1157
0.01 209(25) 5.9(6) 5.8(6) 12(1) 452 240 h samples 50°C
0.001 78(8) 1.6(2) 1.0(1) 2.8(3) 159 0.1 780(142) 139(14) 39(4) 397(40) 2651
DI 45(4) 0.84(8) 0.35(4) 0.66(7) 0.01 510(101) 21(2) 6.9(7) 16(2) 1055
0.001 106(23) 3.4(3) 0.26(3) 0.59(6) 208
5 h samples 50°C DI 72(7) 2.4(2) 0.49(5) 0.70(7)
0.1 777(101) 60(6) 15(1) 95(9) 1808
0.01 242(31) 14(1) 4.1(4) 10(1) 514 240 h samples 90°C
0.001 81(11) 3.3(3) 0.53(5) 1.8(2) 165 0.1 731(133) 190(19) 71(7) 250(25) 2408
DI 49(5) 1.7(2) 0.50(5) 0.67(7) 0.01 459(91) 33(3) 4.0(4) 10(1) 948
0.001 108(24) 5.2(5) 0.14(1) 0.30(3) 213
5 h samples 90°C DI 100(10) 4.8(5) 0.77(8) 1.4(1)
0.1 1000(130) 214(21) 32(3) 392(39) 3118 Note: Concentrations given in parentheses indicate the greatest amount
0.01 327(41) 35(3) 3.9(4) 15(2) 709 of uncertainty in the last digits of the concentration. [For example 239(24)
0.001 89(11) 6.1(6) 0.33(3) 1.5(1) 180 means 239 ± 24 ppm and 4.0(8) means 4.0 ± 0.8 ppm.]
DI 57(6) 2.4(3) 0.50(5) 0.55(5) *Sr values are calculated based on charge balance.
434 PALMER AND GUNTER: ION-EXCHANGE IN CLINOPTILOLITE-RICH ROCKS, PART II

cause our goal was to observe the behavior of the outgoing


cations; in retrospect, we could have measured it in the 0.01
and 0.001 M solutions with proper dilutions.
After our original experiments, we performed other tests to
determine if the solid samples might have been dissolving. First,
we used the same treatment times and conditions but substi-
tuted deionized water for the SrCl solutions to obtain a “blank.”
We also prepared a 1 M solution of SrCl and placed 4 g of
clinoptilolite in it at 90 °C. After 6 h, the pH of the liquid was
6.2 and after 72 h it was 6.0; thus, the liquid was not suffi-
ciently acidic to dissolve the sample. A powder XRD scan of
the solid after 2 weeks of storage was similar to the natural
sample, again suggesting no solid dissolution. Our observa-
tions were confirmed by those of Stolz and Armbruster (1997),
who found that even though the pH dropped to 4 in their Sr
single-crystal exchange experiments, no sample dissolution
occurred.

RESULTS
Solutions with higher original concentrations of Sr were
expected to contain higher concentrations of Na, K, Ca, and
Mg after treatment due to the greater availability of Sr for ex-
change. Results of ICP-AES analyses are given in Table 1. The
results of the analyses can be categorized by the exchange be-
havior of clinoptilolite with monovalent cations and the be-
havior of clinoptilolite with divalent cations.
Monovalent cations exchange out of the mineral and into FIGURE 2. Sodium exchange out of clinoptilolite as a function of
solution, increasing in concentration with time, temperature, time, temperature, and Sr concentration. As Sr concentration increases
and concentration of Sr (Figs. 2 and 3). The Na and K results from 0.001 M to 0.1 M, the amount of Na exchange out of the mineral
and into solution also increases. The amount of Na exchange also
indicate that the majority of monovalent exchange occurs within
increases with temperature. The majority of exchange occurs within
the first 24 h. Some re-exchange of K back into the mineral the first 24 h of the experiment. The Na concentration is lowest in the
occurs in the 0.001 M samples exchanged for 24 h. This re- DI samples.
exchange was observed by Woods and Gunter (2001) for K in
solutions of 1 M CsCl exchanged with 4 g of clinoptilolite. All
three Sr treatments showed greater concentrations than the DI
treatment.
Divalent cations behaved similarly in the 0.1 M samples
(Figs. 4 and 5). However, as Sr availability decreased with lower
Sr concentrations, divalent cations exchanged out of the min-
eral less as temperature increased. The divalent cations conse-
quently exhibited the opposite trend from the monovalent
cations in the 0.01–0.001 M samples. In the 0.001 M samples,
both Ca and Mg exhibited a similar trend in their behavior. The
samples treated at 5 and 21 °C experienced an increase in ex-
change with time, whereas the samples treated at 50 and 90 °C
experienced a decrease in exchange with time. The divergence
that occurred between the high- and low-temperature samples
suggests that some process occurs between 21 and 50 °C that
affects the exchangeability of the divalent cations. However,
the concentrations of the divalent cations in the DI samples

FIGURE 3. Potassium exchange out of clinoptilolite as a function


of time, temperature, and Sr concentration. Exchange of K increases
with increasing Sr concentration and temperature. The mineral
experiences re-exchange of K for 0.001 M samples exchanged for 24
h. The K concentration is lowest in the DI samples.
PALMER AND GUNTER: ION-EXCHANGE IN CLINOPTILOLITE-RICH ROCKS, PART II 435

FIGURE 4. Calcium exchange out of clinoptilolite as a function of FIGURE 5. Magnesium exchange out of clinoptilolite as a function
time, temperature, and Sr concentration. Exchange increases with of time, temperature, and Sr concentration. The 0.001 M samples exhibit
temperature in only the 0.1 M samples and decreases with temperature the divergence between the high and low temperature samples as shown
in the 0.01–0.001 M samples. The 0.001 M samples exhibit a divergence by Ca. The Mg concentration in the DI sample is similar to that in the
between those samples exchanged at 5 and 21 °C and those exchanged 0.001 M sample.
at 50 and 90 °C. The Ca concentration in the DI sample is similar to
that in the 0.001 M sample.

were very similar to those found in the 0.001 M samples. The from solution. However, Figure 6 shows that for the 0.01–0.001
difference between the 0.001 M samples and the DI samples is M solutions, more cations were displaced from the mineral than
that greater exchange occurred at lower temperatures for the there was Sr to replace them. The elevated concentrations of
0.001 M samples and greater exchange occurred at higher tem- cations in these samples may have come from the surface of
peratures for the DI samples. the mineral rather than from inside the framework of the min-
Figure 6 illustrates the calculated exchange of Sr into the eral or from unidentified impurities in the rock.
mineral based on the cations that exchanged out of the mineral
and into solution. The dotted lines represent the amount of Sr DISCUSSION
actually available in the original solution. According to the Although full exchange of cations out of clinoptilolite does
chemical formula of the clinoptilolite used in this study, the not occur in any sample, higher concentrations of Sr promote
maximum concentrations of Na, K, Ca, and Mg that could ex- more exchange of outgoing cations. Others investigators have
change out of 4 g of clinoptilolite into 100 mL of solution were noted the difficulty of exchanging clinoptilolite into a
1237, 701, 195, and 466 ppm respectively. For all of these cat- homoionic form. Semmens and Martin (1988) observed the
ions to exchange out of the mineral, approximately 4846 ppm reluctance of clinoptilolite to release Ca and K upon exchange
Sr would have to exchange into the mineral for charge bal- with Na, and Gunter et al. (1994) only obtained Pb exchange
ance. Full exchange of cations out of the mineral did not occur in one of four natural heulandites. Our results indicate a reluc-
for any of the samples at any concentration of Sr. For each tance of clinoptilolite to release K. Of the 466 ppm Ca avail-
molar concentration of Sr, 8762, 876.2, and 87.62 ppm Sr were able in 4 g of clinoptilolite, 455 ppm Ca (or 98%) exchanged
available for exchange in the solution. Calculations suggest that out of the mineral in a 0.1 M solution after 24 h. In the same
for the 0.1 M solutions, not all of the available Sr was removed sample, however, only 259 ppm (or 37%) K exchanged out of
436 PALMER AND GUNTER: ION-EXCHANGE IN CLINOPTILOLITE-RICH ROCKS, PART II

FIGURE 7. The relationship of cation radius to hydrated cation


radius at 25 °C [from Nightingale (1959)]. Smaller, divalent cations
have larger hydrated radii than larger, monovalent cations.

channel has an aperture of about 3.3 x 4.6 Å, and the C channel


has an aperture of about 2.6 x 4.7 Å (Meier and Olson 1992).
The hydrated radii of Ca2+, Mg2+, and Sr2+ are 4.12, 4.28, and
4.12 Å respectively, which are too large to move about freely
in any of the channels without removal of their associated wa-
ter molecules. The Mg and Ca results suggest that as tempera-
FIGURE 6. Calculated Sr exchange into clinoptilolite based on the ture increases, the radii of the ingoing and outgoing cations
concentrations of outgoing cations in sample solutions. Strontium change to a point where they can no longer move about freely
exchange in clinoptilolite increases with increasing time, temperature, in the channels. Presumably, the hydrated radii of divalent cat-
and Sr concentration. The dotted lines indicate the total amount of Sr ions increase with temperature due to the expansion of water
available in the solution. In the 0.01–0.001 M samples, more cations
with heat. However, cation exchange does occur when the Sr
are released from the mineral than there is Sr available to exchange
into the mineral.
concentration is 0.1 M, suggesting that divalent cation exchange
can occur when the Sr concentration is high enough. However,
when Sr concentration is low, lower temperatures seem to fa-
the mineral of the 701 ppm available. In this sample, 932 ppm vor more exchange of divalent cations.
(or 75%) of the 1237 ppm Na available exchanged out of the Although divalent cations do not exchange out of the min-
mineral and 109 ppm (or 56%) of the 195 ppm Mg available eral easily at high temperatures, overall exchange of Sr into
exchanged out of the mineral. Incidentally, a 0.1 M solution at the mineral is promoted at higher temperatures. This is because
90 °C for 24 h was the most effective environment in this study the samples contained more monovalent than divalent cations.
for achieving maximum exchange of cations out of the Figure 6 shows that a temperature of 90 °C and a time of 24 h
clinoptilolite. are necessary to achieve maximum exchange of clinoptilolite
Greater concentrations of divalent cations exchanged out for most concentrations of Sr.
of the mineral and into solution at lower temperatures in the The results presented in this work and those of Woods and
lower concentration solutions. This behavior could be attrib- Gunter (2001) are largely phenomenological. We have at-
uted to changes in the hydrated radii of the divalent cations. tempted to explain our observations, but with only limited suc-
Assuming that all cations exchange into and out of the mineral cess. However, the significance of both of these studies is that
with their associated water molecules, the radius of the hydra- they show that when a natural clinoptilolite is undergoing ex-
tion sphere is a factor in the exchangeability of a given cation. change in an aqueous solution the outgoing cations behave in
Because of their high charge, divalent cations hold water mol- very different manners. The outgoing cations are affected by
ecules more tightly than monovalent cations; therefore, hydrated the cation in the liquid, Na and Cs in Woods and Gunter (2001),
monovalent cations may be able to adjust their size by shed- and by the concentration of the cations, Sr in the case of this
ding water molecules in order to fit into a site or through a study.
channel. Divalent cations may not shed their water molecules Most of the studies of cation exchange in clinoptilolite start
as easily and therefore may not be able to maneuver easily into with a pretreated sample, usually an Na-exchanged sample,
and out of sites and through channels. Hydrated divalent cat- which is then exchanged with some single or multiphase solu-
ions have larger radii than hydrated monovalent cations, which tion. This type of experiment is useful for many purposes but
is due to their higher charge. Figure 7 illustrates the relation- not to model geological conditions. Natural zeolites will con-
ship between cation radius and hydrated cation radius at 25 °C tain several exchangeable cations in differing amounts. An
(Nightingale 1959). The A channel is the largest channel in important question to answer is how the individual cations be-
clinoptilolite, with an aperture of about 3.0 x 7.6 Å. The B have when they interact with a natural fluid. Of course, the
PALMER AND GUNTER: ION-EXCHANGE IN CLINOPTILOLITE-RICH ROCKS, PART II 437

natural fluid will usually contain more than one element. Thus, Las, T. (1989) Use of natural zeolites for nuclear waste treatment, 273 p. Ph.D.
dissertation, University of Salford, U.K.
what are needed are exchange experiments with natural samples Leppert, D. (1990) Heavy metal sorption with clinoptilolite zeolite: Alternatives for
and multicomponent liquids to truly simulate the natural envi- treating contaminated soil and water. Mining Engineering, 42, 604–608.
ronment. Mason, B. and Sand, L.B. (1960) Clinoptilolite from Patagonia: The relationship
between clinoptilolite and heulandite. American Mineralogist, 45, 341–350.
Meier, W.M. and Olson, D.H. (1992) Atlas of zeolite structure types, 200 p.
ACKNOWLEDGMENTS Butterworth-Heinemann, London.
The authors thank British Nuclear Fuels for partial funding of this project Mumpton, F.A. (1960) Clinoptilolite redefined. American Mineralogist, 45, 351–
and for providing samples of the clinoptilolite-rich tuff used in this study. The 369.
authors also thank Scott A. Wood, Bill Shannon, and Leslie Baker for their Mumpton, F.A. (1978) Natural zeolites: A new industrial mineral commodity. In
assistance in the collection of the ICP-AES data. We also thank our three anony- L.B. Sand and F.A. Mumpton, Eds., Natural zeolites: Occurrence, properties
mous reviewers and David Cole, the associate editor, for suggesting revisions and use, 546 p. Pergamon Press, Oxford.
of this manuscript. Nightingale, E.R., Jr. (1959) Phenomenological theory of ion solvation. Effective
radii of hydrated ions. Journal of Physical Chemistry, 63, 1381–1387.
Palmer, J.L. and Gunter, M.E. (2000) Optical properties of natural and cation-ex-
REFERENCES CITED changed heulandite group zeolites. American Mineralogist, 85, 225–230.
Ames, L.L., Jr. (1960) The cation sieve properties of clinoptilolite. American Min- Schaller, W.T. (1932) The mordenite-ptilolite group; clinoptilolite, a new species.
eralogist, 45, 689–700. Journal Mineralogical Society of America, 17, 128–134.
Ames, L.L., Jr. (1961) Cation sieve properties of the open zeolites chabazite, Semmens, M.J. and Martin, W.P. (1988) The influence of pretreatment on the ca-
mordenite, erionite, and clinoptilolite. American Mineralogist, 46, 1120–1131. pacity and selectivity of clinoptilolite for metal ions. Water Research, 22, 537–
Armbruster, T. and Gunter, M.E. (1991) Stepwise dehydration of heulandite- 542.
clinoptilolite from Succor Creek, Oregon, U.S.A.: A single-crystal X-ray study Semmens, M.J. and Seyfarth, M. (1978) The selectivity of clinoptilolite for certain
at 100 K. American Mineralogist, 76, 1872–1883. heavy metals. In L.B. Sand and F.A. Mumpton, Eds., Natural zeolites: Occur-
Armbruster, T. and Gunter, M.E. (2001) Crystal Structures of Natural Zeolites. In rence, properties and use, 546 p. Pergamon Press, Oxford.
Reviews in Mineralogy and Geochemistry, Mineralogical Society of America, Sheppard, R.A. and Gude, A.J. (1969) Diagenesis of tuffs in the Barstow Forma-
Washington, D.C., in press. tion, Mud Hills, San Bernardino County, California. U.S. Geological Survey
Baxter, S.G. and Berghauser, D.C. (1986) The selection and performance of the Professional Papers, 634, 1–35.
natural zeolite clinoptilolite in British Nuclear Fuels’ Site Ion Exchange Efflu- Stolz, J. and Armbruster, T. (1997) Mg2+-, Mn2+-, Cd2+-, Sr2+-, and Cu2+-exchange in
ent Plant, SIXEP. In R.G. Post, Ed., Waste management volume 2: High level heulandite single-crystals: X-ray structure refinements. 5th International Con-
waste, 625 p. Proceedings on the Symposium on Waste Management, Tucson, ference on the Occurrence, Properties, and Utilization of Natural Zeolites, Is-
AZ. chia, Italy.
Barkatt, A. and Van Konynenburg, R. A. (1993) Thermodynamics of ion-exchange Surdam, R.C. (1981) Zeolites in closed hydrologic systems, 4, 65–91. In Reviews
between Na+/Sr2+ solutions and the zeolite mineral clinoptilolite. Materials Re- in Mineralogy, Mineralogical Society of America, Washington, D.C.
search Society Symposium Proceedings, 333, 731–738. Woods, R.M. and Gunter, M.E. (2001) Na- and Cs-exchange in a clinoptilolite-rich
Boles, J.R. (1972) Composition, optical properties, cell dimensions, and thermal rock: analysis of the outgoing cations in solution. American Mineralogist, 86,
stability of some heulandite group zeolites. American Mineralogist, 57, 1463– 424–430.
1493. Wüst, T., Stolz, J., and Armbruster, T. (1999) Partially dealuminated heulandite pro-
Coombs, D.S., Alberti, A., Armbruster, T., Artioli, G., Colella, C., Galli, E., Grice, duced by acidic REECl3 solution: A chemical and single-crystal X-ray study.
J.D., Liebau, F., Mandarino, J.A., Minato, H., Nickel, E.H., Passaglia, E.,Peacor, American Mineralogist, 84, 1126–1134.
D.R., Quartieri, S., Rinaldi, R., Ross, M., Sheppard, R.A., Tillmanns, E., and Yang, P. and Armbruster, T. (1996) Na, K, Rb, and Cs exchange in heulandite single-
Vezzalini, G. (1997) Recommended nomenclature for zeolite minerals: Report crystals: X-ray structure refinements at 100 K. Journal of Solid State Chemis-
of the Subcommittee on Zeolites of the International Mineralogical Associa- try, 123, 140–149.
tion, Commission on New Minerals and Mineral Names. The Canadian Miner- Yang, P. and Armbruster, T. (1998) X-ray single-crystal structure refinement of NH4-
alogist, 35, 1571–1606. exchanged heulandite at 100 K. European Journal of Mineralogy, 10, 461–471.
Gunter, M.E., Armbruster, T., Kohler, T., and Knowles, C.R. (1994) Crystal struc- Zanetti, K.A. and Gunter, M.E. (1997) Time and temperature dependence of Cs-
ture and optical properties of Na- and Pb-exchanged heulandite-group zeolites. exchange in microcrystalline clinoptilolite. 5th International Conference on the
American Mineralogist, 79, 675–682. Occurrence, Properties, and Utilization of Natural Zeolites, Ischia, Italy.
Hay, R.L. (1981) Geology of zeolites in sedimentary rocks, 4, 53–64. Reviews in
Mineralogy, Mineralogical Society of America, Washington, D.C.
Hey, M.H. (1930) Studies on the zeolites part I: General review. Mineralogical
Magazine, 22, 422–437. MANUSCRIPT RECEIVED NOVEMBER 26, 1999
Hey, M.H. and Bannister, F.A. (1934) Studies on the zeolites part VII: “Clinoptilolite”, MANUSCRIPT ACCEPTED NOVEMBER 25, 2000
a silica-rich variety of heulandite. Mineralogical Magazine, 23, 556–559. MANUSCRIPT HANDLED BY DAVID R. COLE

You might also like