Você está na página 1de 11

Geochimica et Cosmochimica Acta, Vol. 60, No. 2, pp.

185- 195, 1996


Copyright 0 1996 Else&r Science Ltd
Pergamon Primed in the USA. All rights reserved
C016-7037/96$15.00 + .I0

0016-7037(95)00392-4
Solubility of jarosite at 4-35°C
DIRK BARON and CARL D. PALMER
Department of Environmental Science and Engineering, Oregon Graduate Institute of Science
and Technology, Portland, OR 97291- 1000, USA

(Received March 24, 1995; accepted in revised form October 11, 1995)

Abstract-The solubility of jarosite (KFe,(SO,),(OH),) was studied in a series of dissolution experiments.


The experiments were conducted at 4-35°C and at pH values between 1.5 and 3.0 using a synthetic jarosite
with a composition very close to ideal. The solids were left in the reaction vessel for up to 6 months.
Equilibrium was established in the experiments after approximately 3 to 4 months. The log Ksp for the
jarosite dissolution reaction
KFe,(SO&(OH), + 6H’ = K’ + 3Fe7+ + 2SO:- + 6H,O
at 25°C is determined to be - 11.O C 0.3. From this measured solubility product, the free energy of
formation, AC&, is calculated to be -3309.8 + 1.7 kJ mol-‘. Based on the temperature dependence of
the solubility product, the enthalpy of reaction at 25”C, AHF,29X, is -45 & 5 kJ mol-‘, the entropy of
reaction, AS&, is -350 + 40 J mol-’ K-‘, and the heat capacity of the reaction, AC,,,, over the tem-
perature range of the experiments is determined to be -2.1 ? 0.2 kJ mol-’ K-’ The rate of the dissolution
reaction can be described by a first-order model.

1. INTRODUCTION ubility product varying by more than seven orders of magni-


Jarosite (KFe,(SO&(OH),) is a common mineral in acidic, tude. To evaluate the conditions under which jarosite may
sulfate-rich environments formed by the oxidation of sulfides, form and remain stable, more accurate thermodynamic data
especially pyrite. Such environments include acid sulfate soils are needed. Such information is also required to further es-
formed from pyrite-bearing sediments (Van Breemen, 1973; tablish the thermodynamic mixing terms in the K-H30-Na
Carson and Dixon, 1983), weathering of sulfide ore deposits jarosite solid-solution system (Alpers et al., 1989). The pur-
(Bladh, 1982), acid-hypersaline lake sediments (Long et al., pose of this study is to measure the solubility product of ja-
1992; Alpers et al., 1992), acid mine drainage (Nordstrom, rosite at temperatures typical of weathering or supergene con-
1977; Chapman et al., 1983; Alpers et al., 1989), and weath- ditions.
ering of coal refuse from pyritic coal seams (Sullivan and
Sobek, 1982). Jarosite has also been reported in hot springs 2. PREVIOUS STUDIES
(Tkachenko and Zotov, 1974) and hydrothermal environ-
Only a few experimental studies of the soluhility of jarosite have
ments (Keith et al., 1979; Stoffregen and Rye, 1992). Precip- been conducted and we are not aware of any calorimetric studies of
itation of jarosite is also of interest to metallurgists because its free energy of formation. There remains a large degree of uncer-
of its ability to scavenge unwanted elements from hydromet- tainty about the thermodynamic properties of jarosite. If the disso-
allurgical ore processing solutions (Dutrizac, 1983). lution of jarosite is written as
Jarosite is a member of the jarosite-alunite group of KFeJ(SO&(OH), + 6H’ c K+ + 3Fe’+ + 2Sd,- + 6HZ0. (1)
isostructural minerals described by the general formula
then the log of the ion activity product (IAP) is
AB,(SO&(OH),, where the B site is occupied by Fe’+ (jaro-
sites) and Al’+ (alunites) and the A site is occupied most com- log IAP = log (K+) + 3 log (Fe”) + 2 log {So?,- ]
monly by K’, Na’, and HjO+ (hydronium) (Parker, 1962;
+ 6 log {Hz01 + 6 pH, co
Kubisz, 1964; Brophy and Sheridan, 1965). Jarosite is the
potassium-iron endmember of the jarosite-alunite group. Ex- where brackets denote aqueous activity. At equilibrium, the IAP is
tensive substitution by other ions within the jarosite crystal equal to the solubility product, Ksp. The log Kspreported for reaction
1 at 25°C varies from -7.12 to -14.8 and the reported free energy
structure has also been reported (Dutrizac and Kaiman, 1976). of formation, AG;i&, ofjarosite varies from -3317.9 to -3192 -+ 25
Both natural and synthetic jarosites often have a significant kJ mol-’ (Table 1). The reasons for the large variation include (1)
amount of H,O’-substitution on the alkali position (Brophy inconsistent thermodynamic data for calculation of free energies and
and Sheridan, 1965; Kubisz, 1970; Dutrizac and Kaiman, aqueous ion activities, (2) substitution of other ions. particularly hy-
1976). Deficiency in Fe with values for the Fe:S04 molar ratio dronium, in the jarosite crystal structure, and to a smaller degree (3)
variation introduced by different experimental approaches, and (4)
significantly lower than the ideal 3:2 stoichiometry is also analytical uncertainty.
common (Kubisz, 1970; Alpers et al., 1989). Values of the Brown (1970) conducted a dissolution experiment in which syn-
Fe:SO, molar ratio as low as 2.33:2 (Ripmeester et al., 1986) thetic jarosite was dissolved at 25°C for 6 months and a free-drift
and 2.20:2 to 2.57:2 (Htiig et a1.,1984) are reported. precipitation experiment from which jarosite was precipitated from
a supersaturatedsolution. In neither of the experiments was equilib-
Despite the common occurrence of jarosite and its interest rium demonstrated. The free energy of formation AG& was cal-
to geochemists, geologists, and metallurgists, uncertainty ex- culated as -3192 + 25 kJ mol-’ from the dissolution experiment and
ists regarding the solubility of jarosite with the reported sol- as -3276 5 84 from the precipitation experiment. Zotov et al. (1973)
185
186 D. Baron and C. D. Palmer

Table 1. Summary of solubilities and AGo,,,,, values reported for jarosite.

Reported Reported Comments Recalculated


log I<sp at 25°C AGo, AGo,
(kJ mol”) (Id mol.‘) *
ALLISON et al. -14.8 -333 1.4
(1990)
ALPERS et al. -11.14 -3300.2k2.6 for (Kt,s,Nq,H,Ott, ,J-jarosite -3300.1i2.6
(1989) precipitated from acid mine
drainage water in the laboratory at
ambient temperature
BLADH (1982) -7.12 -3287.6
BROWN (1970) -3216&t from precipitation experiment
-3192*25 from dissolution experiment

-3302G4 recalculated by ZOTOV et al. -3302i84


-33OOi25 (1973) who noted an arithmetic -33OOzt2S
error in Brown’s calculations and
used thermodynamic data from
NAUMOV et al. (1971)

-3299 recalculated from the dissolution


experiment by VAN BREEMEN
(1973) using thermodynamic data
from ROBIE and WALDBAUM
( I9681
CHAPMAN et al. -9.21 basedon KASHKAY et al. (1975)
(1983) and thermodynamic data from
NAUMOV et al. (1971)
HLADKY and -9.08 based on KASHKAY et al. (1975)
SLANSKY (19811
KASHKAY et al. -3299.7+4 ‘free-drift’ precipitation -3299.6*4
(1975) -15.8 (at 100°C) -3 184 (at I OO”C1 exoeriments
STOFFREGEN -3416.3il.7 at 2oO”C, 100 bar
(1993) (200°C. 100 bar)
VLEK et al. (1974) -14.56 -3310.4 used a ‘chelation method’ -3331.4
ZOTOV et al. -33058i4 basedon a jarosite precipitated -3305.7
(1973) from a natural water sample

-3317.9 estimated from a natural (Na,K)- -3317.8


jarosite precipitated at 45°C
present study -1l.OztO.3 -3309.8t1.7

’ AGor,2a8Wo recalculated using the free energies for ions used in the present study (Table 2).

noted an arithmetical error in the calculation of Brown (1970) and AG;an.,oo barof -3416.3 ? 1.7 k.l mol-‘. Chapman et al. (l983),
recalculated AGo,,29nas -3300 + 25 kJ mol-’ from the dissolution Bladh (1982). Van Breemen (1973) and Hladky and Slansky (1981)
experiment and -3302 2 84 k.I mall from the precipitation exper- have calculated solubility or free energy of formation based on re-
iment. Vlek et al. (1974) dissolved a natural jarosite in a solution interpretation of the above experimental studies.
containing the chelating agent EDTA to increase total Fe(M) con- One important source of the variation in the values for the free
centrations and calculated a AC&a of -3310.4 kJ mall’ and a log energy of formation of jarosite is the use of different values for the
Ksp of -14.56. Kasbkay et al. (1975) conducted three precipitation free energies for the ions used by different researchers. To compare
experiments in which jarosite was allowed to precipitate from super- the reported values better, we have recalculated the free energies for
saturated solutions for 15 months at 25 ? 5°C and calculated the formation of jarosite using the free energies for the ions used in
AG&s as -3299.7 + 4 kJ mol-‘. From another precipitation exper- this study and listed in Table 2. The results of these calculations are
iment conducted at lOO”C, these researchers determined a log KsP.171 included in Table 1. However, even after eliminating the variation
of -15.83 and calculated AC;,,, as -3184 k.l mall’. Zotov et al. due to different free energies of the ions, a large variation in the
(I 973) present two estimates of the free energy of formation of jaro- reported free energies of formation of jarosite remains.
site. Based on the synthesis of (KasNq.,)Fe,(SO,),(OH), in a natural The values for the free energy of formation and the solubility prod-
thermal water at 45°C they estimate AC;>, as -3317.9 k.l mall’, uct of jarosite that are currently considered the most reliable (e.g..
and based on a precipitate from a natural water sample, they estimate Alpers et al., 1989; Nordstrom and Munoz, 1994) are the AG&
AG;ms as -3305.8 2 4 k.l mall’. Stoffregen (1993) measured the = -3299.7 2 4 k.l mol-’ repotted by Kashkay et al. (1975) and the
solubility of jarosite under hydrothermal conditions and calculated a log Ksp = -9.21 calculated from this free energy by Chapman et al.
Solubility of jarosite 187

Table 2. Thermodynamic data used in calculations.

Formula State AG“, AH’, log I<I Source


(kJ moP) (kJ mol”)
Fe’+ aq -17.87kl.O - 2
FeSO,’ aq 16.36 3.92 1
FeHSO: aq - 2.48 4
Fe(S0,); aq - 19.3 5.42 1
FeOH’+ 4 43.5 -2.19 1
Fe(OH),+ aq - 71.6 -5.67 5
Fe&OH),4 aq 56.5 -2.95 1
K’ aq -282.5&o. 1 - 3
KSO, aq 9.41 0.85 1
so,2- aq -744.oko.4 - 3
HSO, aq -22.4 1.98 3
OH- aq 55.836 -13.998 1
W 1 -237.14ti.04 - 3

Sources: (1) ALLISON et al. (1990)


(2) NAUMOV et al. (1971)
(3) COX et al. (1989)
(4) BALL et al. (1987)
(5) NORDSTROM and MUNOZ (1994)

(1983). These values are considered the most reliable because they spectroscopy (SEM/EDX), Fourier Transform Infrared Spectroscopy
are based on actual solubility measurements and a consistent ther- (FTIR), and thermogravimetric analysis (TGA). A small amount of
modynamic database. However. Alpcrs et al. (1989) noted that the precipitate was digested in HCI and analyzed for K and Fe using
free energies of formation of K-HSO-Na jarosite solid solutions are atomic absorption spectroscopy (AAS) and SO, using high perfor-
more consistent with the free energy for pure jarosite reported by mance ion chromatography (HPIC).
Vlek et al. (1974) and Zotov et al. (1973), indicating a lower free
energy of formation and lower solubility than the Kashkay et al.
(1975) value. It is not surprising, therefore, that Alpers et al. (1989) 3.3. Dissolution Experiments
felt that additional experimental work was needed to determine ac- Two sets of dissolution experiments were implemented. The first
curately the solubility and free energy of formation of jarosite. was conducted at 25°C and the initial pH varied between I .5 and 3.0.
In the second set of experiments, the temperature was varied between
3. EXPERIMENTAL METHODS 4 and 35°C with an initial pH of 2.0 for all experiments. For the
dissolution experiments, 5 to 30 mg of the synthetic jarosite were
3.1. Synthesis of Jarosite
added to ultrapure water with the pH adjusted to the desired value
Synthetic jarosite often has a K deficiency because of hydronium using HCIO,. To avoid formation of ferric oxyhydroxides at the
substitution in the alkali ion position (Kubisz, 1970; Dutrizac and higher pH values, 6.3 x 10m4and 3.2 X lo-” M of CaS04.2H20
Kaiman, 1976). The factors that control hydronium substitution in- were added to the experiments with initial pH 2.6 and 3.0, respec-
clude the KC activity and the pH. To minimize hydronium substitt- tively. The solutions were placed in 20 mL glass vials and stirred
tion and to obtain a synthetic product with a composition close to the with a stirbar at a moderate rate (about 100 rpm) to provide good
ideal formula, K* was added in excess to the synthesis solution as mixing. The temperature was maintained to within O.l”C of the de-
KOH, thus, simultaneously increasing the activity of K+ and lowering sired value using circulating water baths. The experiment with initial
the H,O+ activity. Jarosite was prepared by dissolving 5.6 g of rea- pH 2.0 and at 25°C (KJAR-2.0) was conducted prior to the other
gent grade KOH and 17.2 g of reagent grade Fe2(S04),.5HZ0 in 100 experiments to establish the approximate time required to achieve
mL H,O at 95”C, 1 atm. The solution was placed in a covered beaker equilibrium. To allow more frequent sampling and analysis for K,.t,
on a hot plate and continuously stirred. After 4 h, the precipitate was Fe,,,,, SO&, , and pH for every sample, this experiment was conducted
allowed to settle and the residual solution decanted. The precipitate with 100 mg of synthetic jarosite in 500 mL of ultrapure water in a
was then washed thoroughly with ultrapure water (18 megaohm polyethylene bottle. The starting conditions for ail dissolution ex-
cm-‘) and dried at 110°C for 24 h. periments are listed in Table 3. All experiments were conducted in
triplicate. The experiments were sampled over time to determine
3.2. Characterization of Synthetic Jarosite when equilibrium was achieved. Ten samples were collected from
the experiments in which the pH was varied and nine samples were
The synthetic solid was characterized using powder X-ray diffrac- collected from the experiments in which the temperature was varied.
tion (XRD), scanning electron microscopy with energy dispersive For each sample, 1 mL of the jarosite suspension was withdrawn.
D. Baron and C. D. Palmer

Table 3. Initial experimental conditions.

amount of amount initial Temperature [CaSOg 2H,O] Duration of


solution of solid pH (“Cl (mol L-‘) Experiment
(mL) (w) (days)
KJAR-1.5 20 30 1.50 25 0 166
KJAR-2.0 500 100 2.00 25 0 149
KJAR-2.3 20 5 2.30 25 0 166
KJAR-2.6 20 5 2.62 25 6.30 x IO4 166
KJAR3.0 20 5 3.02 25 3.16 x lo” 166
KJAR4C 20 5 2.00 4 0 176
KJAR-15C 20 5 2.00 15 0 176
KJAR-3% 20 5 2.00 35 0 176

All experiments were conducted in triplicate.

The samples were filtered using a 0.1 pm filter to remove suspended Thermogravimetric analysis was conducted by heating the
solids and then analyzed for potassium using AAS. After the K con- synthetic jarosite from 30 to 900°C at a rate of 20”Umin (Fig.
centration had not changed significantly (?5%) for at least three can-
secutive samples, a 4 mL sample was taken and analyzed for pH, 2). Curve A shows the weight loss vs. temperature and curve
Fe,,, and K,,, using AAS and SO&, using HPIC. Since the dissolution B its derivative. The total weight loss over the interval is
experiments were conducted in oxidizing HCIO,, solutions and no 38.5%. The curves show that the weight loss occurs in three
reductants capable of reducing ferric iron were present, it was as- principal temperature intervals: (1) a 2.5% weight loss be-
sumed that all Fe was present as ferric iron.
tween 200 and 320°C represented by a peak at about 260°C.
(2) an 11% weight loss in the interval between 350 and 450°C
4. EXPERIMENTAL RESULTS represented by a peak at about 430°C and (3) a 25% weight
loss between 560 and 800°C. represented by a group of four
4.1. Solid Characterization peaks at about 620, 720,760, and 785°C.
The weight loss in the interval from 200 to 320°C is com-
The yellow precipitate produced in the synthesis was iden- monly attributed to the loss of hydronium (Brophy and Sher-
tified as endmember jarosite by comparing powder X-ray dif- idan, 1965; Kubisz, 1970; Alpers et al., 1989). However, the
fraction patterns with those for jarosite reported in JCPDS weight loss of 2.5% observed in our experiments would cor-
card 22-827 (JCPDS, 1991) (Table 4, Fig. 1). All the peaks respond to about 0.66 moles of H,O’ per mole of jarosite, far
produced by the precipitate could be identified as jarosite exceeding the 0.02 moles of hydronium expected from the
peaks. The absence of unidentified peaks indicates that no slight K deficiency determined in the chemical analysis. Sim-
other crystalline phases are present in the precipitate at de- ilarly, the weight loss of around 2 to 3% in low-temperature,
tectable levels. The unit cell dimensions were calculated as u0 K-rich jarosites, commonly observed and attributed to hydro-
= 7.291 ? 0.005 A and c0 = 17.136 ? 0.015 A. These values nium (e.g., Brophy and Sheridan, 1965; Alpers et al., 1989)
are consistent with the those reported for pure endmember is inconsistent with the reported K+ content of the jarosites.
jarosite with no hydronium substitution (Alpers et al., 1989). It appears that most of the weight loss between 200 and 320°C
Hydronium-jarosite has a,, in the range of 7.34 to 7.36 A and must be from another source. An alternative explanation for
jarosites with appreciable hydronium in solid solution have this weight loss has been presented by Hartig et al. (1984).
an elongated a-axis compared to pure endmember jarosite (Al- They noted that the charge deficiency caused by the Fe7’ def-
pers et al., 1989). Examination with SEM/EDX showed that icit is likely balanced by a partial substitution of HZ0 for OH-,
the precipitate consists of multicrystalline particles with uni- and argue that this water contributes a major portion of the
form concentrations of K, Fe, and S, ranging in size from 10 weight loss between 200 and 320°C. Similar arguments are
to 150 pm. No other crystalline or amorphous phases were given by Kubisz (1972). The synthetic jarosite used in this
observed. study has an Fe deficit of 0.21 moles per mole of jarosite or
Wet chemical analysis yielded a composition of the precip- a resultant charge deficiency of 0.63 positive charges per mole
itate very close that of “ideal” jarosite with a K:Fe:SO, ratio of jarosite. Thus, 0.63 moles of HZ0 would have to be sub-
of 0.98:2.79:2 compared to the ideal 1:3:2 ratio. The measured stituted for OH- to balance the charge deficiency. This cor-
stoichiometry indicates that the synthetic jarosite has almost responds almost exactly to the 2.5% weight loss equivalent to
no hydronium substitution and may have an Fe deficiency of a loss of 0.66 moles of water. The observed weight loss in the
approximately 7%. This deficiency is slightly larger than the interval from 200 to 320°C can therefore almost completely
analytical error of approximately 5%. be attributed to a loss of H,O substituting for OH- to balance
Solubility of jarosite 189

Table 4. Powder x-ray diffraction peaks from synthetic jarosite used in dissolution experiments.

Synthetic jarosite JCPDS Card 22-827


used in this study (JCPDS, 1991) hkl
d-spacing relative d-spacing relative
(4 Intensity (A) Intensity

5.94 30 5.93 45 101


5.72 15 5.72 25 003
5.09 58 5.09 70 012
3.66 22 3.65 40 110
3.55 5 3.55 4 104
3.11 85 3.11 75 021
3.08 100 3.08 100 113
2.969 11 2.965 15 202
2.855 16 2.861 30 006
2.542 21 2.542 30 024
2.302 10 2.302 12 122
2.282 29 2.287 40 107
1.979 37 1.977 45 033
1.935 7 1.937 10 027
1.909 6 1.909 8 009
1.828 30 1.825 45 220
1.769 6 1.776 6 208
1.740 6 1.738 6 223
1.720 5 1.717 6 312
1.690 4 1.690 2 119
1.659 5 1.656 2 1 0 10
1.625 5 1.621 6 134
1.592 5 1.595 6 128
1.572 5 1.572 4 401
1.560 6 1.560 6 3 15
1.551 7 1.552 6 042
1.534 15 1.536 20 226

the charge deficit due to the slight Fe deficiency and does not tification of hydronium in jarosites by FlXX is generally dif-
indicate a significant hydronium substitution. A balanced, ficult and often not conclusive (Wilkins et al., 1974; Rip-
complete formula for the synthetic jarosite is K,&H,O)o.Oz meesteret al., 1986). The only easily identifiable peak asso-
X Fe2.~~(SO~)~(OH)~.~,(H20)o.63. ciated with hydronium in jarosite lies in the 1535- 1575 cm-’
The weight loss in the interval between 350 and 450°C region as a weak peak on the shoulder of a stronger peak at
correspondsto the decomposition of jarosite into KFe(SO& 1635- 1640 cm-‘, associatedwith HZ0 (Kubisz, 1972; Wil-
and Fe203and the loss of the remaining water in the crystal kins et al., 1974). This hydronium peak is absentin the spec-
structure (Kulp and Adler, 1950). The 2.7 moles of water trum of our precipitate.
remaining in the crystal structure would correspond to 10% Based on these analyses, the synthetic jarosite has a com-
by weight which is close to the observed weight loss of position of very close to the ideal stoichiometry with a small
11.O%.The final weight loss between 560 and 800°C repre- deficit of Fe’+ and somerelated substitution of Hz0 for OH-,
sents the seconddecomposition of the KFe(SO& into K2S04 but it does not have appreciablehydronium substitution in the
and FeZOJand the release of 1.5 moles of SO, per mole of alkali position.
jarosite. The observed 25% weight loss is very close to the
expected weight loss of 24%. 4.2. Dissolution Experiments
Comparison of the FTIR scanof the precipitate (Fig. 3) and
reported scansforjarosite (Kubisz, 1972; Wilkins et al., 1974; In the dissolution experiments, the bulk of the reaction oc-
Powers et al., 1975) confirmed the identification as jarosite curred within the first week of the experiment with rates de-
and the absenceof other phases in the precipitate. The iden- clining with time. The evolution of the solution composition
190 D. Baron and C. D. Palmer

3.08 d-spacing (A)


=
c 60 -

rr
10 20 30 40 50 6 0 I”““““““”
4000 3ooa 2ooo 1000 400
2 Theta (degrees) Wave Number (cm-‘)
FIG. 1. Powder X-ray diffraction spectrum of the synthetic jarosite.
FIG. 3. FUR spectrum of the synthetic jarosite. The main vibra-
Some of the strongest peaks and the corresponding d-spacings are
tional bands in the spectrum are marked. They correspond to (Powers
indicated.
et al., 1975) the O-H stretch (3386 cm-‘), HOH deformation (1636
cm-‘), the V~ mode of SO:- (1187 and 1087 cm-‘), OH deformation
(1005 cm-.‘), the uq mode of SOa- (668 and 628 cm-‘), and to the
over time in the dissolution experiment at 25”C, initial pH 2 vibrations of the FeO, coordination octahedron (575, 513, and 477
(KJAR-2.0) is shown in Fig. 4 as an example for the disso- cm-’ ).
lution process. Equilibrium was attained in the dissolution
experiments after 80 to 125 days (Table 5). The equilibrium
are presented in Table 6. The average charge balance error in
compositions of the solutions in the dissolution experiments
the MINTEQA2 calculations was 3%. The error associated
are summarized in Table 5. Based on the measured equilib-
with the log IAP values calculated from the standard deviation
rium pH and K,,, , Fe,,, , and SO&, concentrations, equilibrium
of the triplicate experiments range from 0.13 to 0.24 log units.
aqueous activities of K+. F$+, and SOi- were calculated us-
The analytical error calculated from the precision of the an-
ing the geochemical speciation model MINTEQA2 (Allison
alytical measurements (210%) and the precision of the pH
et al., 1990). Activity corrections were made using the Davies
buffer solutions (20.02 pH units) is 0.25 pH units (assuming
Equation. The MINTEQA2 thermodynamic database was
that the covariance between these parameters equals zero).
modified to include the FeHSOZ$,, ion pair. Other ion pairs
The error calculated from the triplicate experiments is smaller
included in the calculations and thermodynamic data used are
than the analytical error. The error associated with the log
listed in Table 2. Based on the calculated activities of H’, K’,
IAP values reported in Table 6 and used for subsequent cal-
Fe7’, and SO:-, the equilibrium ion activity products (IAP)
culations is the analytical error of 50.25 log units.
were calculated using Eqn. 2. The results of these calculations
As previously discussed, the synthetic jarosite has a com-
position slightly different from the ideal composition on
which Eqn. 2 is based. However, using Eqn. 2 with the ideal
100
stoichiometry is consistent with previous work on the solu-
bility of jarosite and appears justified since the nonideality is
090.
5
!2 0.6 I I I
initial pH 2,25’C
Q66.
E
P 8
8 Fe tot
g 70
8
8 so
4 tot

“Ul 100 300 506


,v , 1
700 OO
6 l,W
A

I 1
mJ(J
A

Time (hours)
K tot

I
3,000
A

4,000

Temperature (“C)
FIG. 4. Concentrations of K,,, (A), Fe,,, (W), and SO& (e) in the
FIG. 2. Thermogravimetric
analysis of the synthetic jarosite. Curve dissolution experiment with initial pH 2.0, equilibrium pH 2.10. Data
A shows the weight loss vs. temperature. curve B is its derivative. points represent the average of triplicate experiments.
Solubility of jarosite 191

Table 5. Final concentrations in the dissolution experiments.

PH W,=l, WI, [W,, Equilibrium Duration of


(mm01 L-l)’ (mmol L’)’ (mm01 L-l)’ established Experiment
after (days) (days)
KJAR-1.5 1.6OkO.03 2.571to.06 1.23dM6 3.37dzO.15 125 166

KJAR3.0 2.10iO.M 0.332dM4 0.178dHM4 0.434iO.005 80 149

KJAR-2.3 2.34~1~0.03 0.313kO.013 0.159dmO7 0.171iO0.018 loo 166

KJAR-2.6 2.6OM.03 0.632iO.06 0.111i00.004 0.05 1Hmo4 120 166

KJAR-3.0 2.986iO.03 3.26ti.28 0.085dI.022 0.013dUlO1 120 166

KJAR4C 2.01iO.02 0.339iO.016 0.217dMO4 0.519~0.016 125 176

KJAR-1SC 2.03ti.02 0.347&0.015 0.225ti.004 0.53OdI.012 100 176

KJAR-3SC 2.01 MO.03 0.528iO.032 0.318iO.019 0.464M.005 100 176

The reported concentrations representthe mean f the standarddeviation from triplicate experiments.

only slight. If one was to use the actual composition as de- varies from - 10.8 1 to - 11.36 (Table 6). A plot of log IAP
termined by the chemical analysis, calculated ion activity vs. pH (Fig. 5) suggests that there is no trend in the data and
products would be approximately 0.5 log units lower than the a t-test indicates that the slope is not significantly different
ones calculated using the ideal stoichiometry. This difference from zero at the 95% significance level (1 = -0.014,3 degrees
is due primarily to substitution of Hz0 for OH- groups. of freedom). The average equilibrium log IAP represents the
To evaluate the effect of different activity correction mod- log Ksp at 25°C and it is calculated as - 11.0 2 0.3. The error
els, ion activity products were also calculated using the ex- in the log Ksp value represents the total standard deviation
tended Debye-Htickel equation. At the low ionic strengths of over all experiments. An F-test indicates that the variance
the experiments, the difference between these two methods is among the log IAP values at different pH values is not sig-
small. For the experiment with the highest ionic strength (0.04 nificantly different from the variance within the triplicate ex-
M), the IAP calculated with the extended Debye-Hiickel equa- periments at the 95% confidence level (F = 2.6, n, = 5, n2
tion is about 0.2 log units greater than the one calculated using = 15, dfi = 4, df2 = 10). At equilibrium. the Gibbs free energy
the Davies Equation. For all other experiments, the difference of reaction at 25°C is given by
between the IAP calculated with the two methods is less than AG&,i”. = AG;‘&K+) + 3AG&(Fe?+)
0.05 log units.
5. DISCUSSION + ‘UG;mW-) + 6~G;mAH@~ - AGm.,aros,te (3)
The log IAP calculated using equilibrium activities from with the free energy of reaction related to the Ksp by
the experiments with different initial pH values and at 25°C AG~~,,,i,,, = -RT In Ksp. (4)

Table 6. Calculated equilibrium activities.

PI-I 1% log log Ionic calculated


w,2-1 IK’I { Fe3+} Strength log IAP
UW
KJAR-1.5 1.60 -3.54 -3.00 -3.44 4.69x lO-2 -10.8lti.25
KJAR-2.0 2.10 -3.99 -3.80 -4.06 1.18x10-’ -11.36fi.25
KJAR-2.3 2.34 -3.84 -3.84 -4.51 6.15~10” -11.03ti.25
KJAR-2.6 2.60 -3.45 -3.99 -5.25 5.08~10” -11.06ti.25
KJAR-3.0 2.99 -2.79 -4.12 -6.37 1.16~10.~ -10.89ti.25
KJAR-4C 2.01 -3.94 -3.71 -3.85 1.34x 1o-2 - 11.06M.25
KJAR-15C 2.03 -3.98 -3.70 -3.89 1.34x1o-2 -11.14ti.25
KJAR-35C 2.01 -3.87 -3.55 -4.14 1.31x1o-2 - 11.68d.25
192 D. Baron and C. D. Palmer

between ( (H30)0,75Na&-jarosite and jarosite, one can fit an


ideal solid solution model to the free energies and estimate
AG&g.larocia from

+ (X, x RT X In (X,) + (1 - XK)


x RTX ln(1 -X,), (6)

where X, is the mole fraction of potassium in the alkali po-


sition (Fig. 6). The AGtq298,jarubltc
was estimated as -3308.4
2 2.2 kJ mol-’ which is in excellent agreement with the value
we obtained in our dissolution experiments.
FIG. 5. Calculated log ion activity product from five dissolution
Key questions that need to be addressed are the reasons for
experiments at 25°C. The solid line represents the average log IAP the apparent discrepancy between the results of this study and
from the five experiments (- 11.O -t 0.3). the results of Kashkay et al. (1975) and which results best
represent the thermodynamic properties of jarosite. Possible
sources of the discrepancy include the thermodynamic data
Solving for the free energy of formation of jarosite and the speciation model used in the two studies. However,
= AG&,(K’)
AG;1298.,arasite + 3AGE&Fe”) the free energies of formation of the ions used in the two
studies are consistent and recalculating the log Kspfrom the
+ 2AG&(SOi-) + 6AGtq&(HZO) + RT In Ksp (3 equilibrium solution compositions given by Kashkay et al.
(1975) with the speciation model used in this study yields a
and using the free energies given in Table 2, we calculate
value of log Ksp= -8.4 ? 0.7, which is even greater than
AGtq*98,jarosite= -3309.8 + 1.7 W mol-’ where the variation
the value calculated by Chapman et al. (1983). Another pos-
represents the error introduced by the uncertainty in the Ksp
sible reason for the apparent inconsistency is the substitution
value only. The error in the free energies of the individual
of hydronium in the crystal structure. Hydronium jarosite,
ions (Table 2) adds an additional uncertainty of +- 4.5 kJ
which has a reported log Ksp of -5.39 and a AG& of
mol-‘. The error introduced by the uncertainty in the free
-3232.1 kJ mol-’ (Kashkay et al., 1975; Chapman et al.,
energies of the individual ions is large compared to the error
1983), is significantly more soluble than jarosite. Hydronium
of the solubility determination. The error in the free energies
substitution in the jarosite crystal structure would therefore
of some of the individual ions may be larger than reported in
lead to a greater overall solubility and a greater AGE,,. A
Table 2, resulting in an even larger uncertainty in the estimate
15-30% hydronium substitution for K, a range commonly
of AG” f.29s.,arosi,e.
For example, for AG&(Fe”) values as
reported for both synthetic and natural jarosite (e.g., Brophy
high as -4.6 kJ mol-’ have been reported (Wagman et al.,
1969). Propagating an uncertainty of 13.3 kJ mol-’ for
AGE&Fe”) through Eqn. 5 results in an overall uncertainty
of +- 43 kJ mol-’ for AG~298,jamsitc. -3220
<
The log Ksp= - 11.O 2 0.3 and the free energy of forma- Ideal Solid Solution
tion, AGlqzg8 = -3309.8 2 1.7 kJ mol-’ are well within the
range of values reported in the literature (Table 1). They are,
however, significantly lower than the commonly accepted val-
ues of log Ksp= -9.21 and AGlqlg8 = -3299.7 rf: 4 kJ mol-’
(Kashkay et al., 197.5; Chapman et al., 1983). In a study of ALLISON et al. (1990)
the solubility of K-Na-H,O-jarosite solid solutions precipi- tic-= - 0 ZOTOV et al. (1973)
a
tated from natural water samples, Alpers et al. (1989) calcu- + BIADH (1978)
-3320 - 0 VLEK et al. (1974)
lated the free energy of formation of these solid solutions and
n present study
noted that their results were not consistent with the free energy
-3340
of formation for jarosite reported by Kashkay et al. (1975). 0.0 0.2 0.4 0.6 1.0

Based on their results, they suggested that the AG;,,, for ja- ((H 3 O),,6Na,6) - Jarosite mok fraction Of K Jarosite
rosite may be closer to the value of -33 10.7 kJ mol-’ reported
FIG. 6. Free energy of formation for K-Na-H,O-jarosite solid so-
by Vlek et al. (1974) or the -3305.8 2 4 kJ mol-’ reported
lutions from Alpers et al. (1989; 0) and Zotov et al. (1973; 0). The
by Zotov et al. (1973). Such values of AG& imply that the solid line represents the fit of an ideal solid solution mixing curve
KSPmay be 1 to 2 log units lower than the commonly quoted betweenjarosite and ((H?O), ,sNan2S)-jarosite.Reported free energies
-9.21. Our results are consistent with the results of Alpers et energies of formation of jarosite from Kashkay et al. (1975; 0) and
al. (1989). We have estimated the jarosite free energy of for- Zotov et al. (1973; 0). and calculated free energies of formation from
Allison et al. (1990; A), Vlek et al. (1974; 0), Bladh (1978; +), and
mation using the free energies for K-Na-H30-jarosite solid from this study (W) are also included for comparison. For consis-
solutions from Alpers et al. (1989) and from Zotov et al. tency, all free energies have been recalculated using the free energies
(1973). Assuming an ideal solid solution in the pseudo-binary of ions given in Table 2.
Solubility of jarosite 193

and Sheridan, 1965; Kubisz, 1970), would result in a solubil- (Nordstrom and Munoz, 1994) with [K+) = lo-“ and
ity product about one to two log units greater than that of pure (SO:-) = lo-’ and taking log Kspof Fe(OH),,,, as +4.89
jarosite. (Allison et al., 1990), we calculate pH, = 5.89 vs. a pH,
Kashkay et al. (1975) report that the compositions of their = 5.29 calculated from Kashkay et al. (1975) and Chapman
precipitates were close to ideal jarosite and that they did not et al. (1983). The solubility product for Fe(OHX(,, in the MIN-
detect any significant hydronium. However, if a hydronium- TEQA2 database (Allison et al., 1990) represents a freshly
enriched jarosite was precipitated as a thin surface layer, it precipitated, poorly crystalline ferric oxyhydroxide (Lang-
may not have been detected in the analysis of the bulk pre- muir and Whittemore, 1971). The solubility product of aged,
cipitate. Solutions would equilibrate with this hydronium-en- more crystalline ferric oxyhydroxide is up to 6 orders of mag-
riched surface layer, rather than with the bulk solid and would nitude lower (Langmuir and Whittemore, 1971) and its sta-
yield greater apparent solubility products for jarosite. bility field would be correspondingly larger. The transition
The experiments conducted by Kashkay et al. (1975) were pH between jarosite and aged, more crystalline ferric oxy-
“free-drift” precipitation experiments in which jarosite was hydroxide would therefore be correspondingly lower.
precipitated from supersaturated solutions. The solution com- As previously discussed, incorporation of the nonideal
position was not held constant during the experiments and stoichiometry of jarosite with a measured Fe-deficit of 0.21
concentrations of K’, Fe(III), and SO:-, as well as pH, were moles of Fe per mole of jarosite in the solubility expression
allowed to drift freely as the precipitation proceeded. In all leads to a calculated log Ksp = - 11.5, about 0.5 log units
three experiments, the solutions at equilibrium were depleted lower than the one calculated using the ideal stoichiometry.
in K relative to Fe and SOi- and enriched in H,O’ compared Using the specific stoichiometry of our jarosite, the transfor-
to the initial solutions. Specifically, the K:SO, ratio at equi- mation to Fe(OH)X,,, is written as
librium decreased by more than an order of magnitude and
the pH decreased by 0.2 to 0.4 units compared to the begin- KFe2.,s(S04)2(OH)s.~,(H20kh~ + 2.37&O
ning of the experiments. It appears likely that the solids pre-
it 2.79Fe(OH& + K’ + 2SO:- + 3H+. (10)
cipitated on the surface when the solutions were close to equi-
librium would be hydronium enriched resulting in a more sol- At equilibrium, the transformation is expressed as
uble surface layer on the bulk jarosite precipitate. The
formation of a hydronium-enriched surface coating is unlikely - 2.79 log
log KsP.,arns,tc Ks~.F~(cJH,,,~
= 1% {K*)
when dissolving the less soluble endmember of a solid solu-
tion as in the experiments reported in this study. We therefore + 2 log (SO:-) + 3 log (H+) - 2.37 log [H,O) (11)
believe that our values better represent the solubility of pure
jarosite. and the transition pH is
The solubility product for jarosite that was obtained in this
study is lower than the commonly cited value of -9.21 (Kash-- + (2.79/3) log
pH, = -‘6 log Ks~.,ctros,,r ~&P.woH,,,,,

kay et al., 1975; Chapman et al., 1983), which suggests that + ‘/3 log (K’ ) + */?log (SO:- )
the stability field of jarosite is larger than previously expected.
The stability of jarosite is generally limited to a relatively - (2.37/3) log (H,O). (12)
narrow range of acidic conditions. As pH increases, jarosite
is transformed to ferric oxyhydroxide, represented in the fol- Substituting our log KSP,,aros,tr = -11.5 calculated for the
lowing discussion as Fe(OH),,,,. The reaction between these non-ideal stoichiometry into Eqn. 12 and using log Ksp
two phases is of Fe(OHh = +4.89 (Allison et al., 1990) yields a pHT
which is 0.17 pH units lower than the value obtained using the
~e3WU40H)6 + 3H20 ideal composition. The transition pH determined using the
= 3Fe(OH),,,, + K’ + 2SO:- + 3H’. (7) log KSP,jama,,ecalculated taking the nonideal stoichiometry into
account is 0.43 pH units higher than the value calculated using
At equilibrium, this transformation can be expressed as the log KSP.jarosia from Kashkay et al. (1975) and Chapman et
al. (1983). For our example water ((K’) = 10m4, (SO:-)
1% Ks~.,awtc - 3 1% &P.~w~H~~,,~ = 1% iK+ 1
= lo-“) we calculate pHr = 5.72 using the ideal stoichiometry
+ 2 log [SO:-) + 3 log {H’) - 3 log (H20). (8) and our log &P.p.rosiucompared to a pHT = 5.89 calculated
using the ideal stoichiometry and a pHT = 5.29 calculated
Therefore, the transition pH, pH,, is with the log KsP.,~~~~,~~ from Kashkay et al. (1975) and Chap-
man et al. (1983).
PHT = ~‘4 log KsP.jarosw + log &P,F~‘~(oH),,,, + ‘13 log 1 K’ 1
The calculated log IAP decreases with increasing temper-
+ ‘1~log (SO:-) - log { H,O) . (9) ature, varying from - 11.06 at 4°C to - 11.68 at 35°C (Table
6, Fig. 7), indicating a negative enthalpy of reaction, AH:,
From Eqn. 9, it becomes apparent that pHT is a function of for the dissolution of jarosite. The dependence of log IAP on
the solubility products of jarosite and Fe(OH),,,,, and the ac- the inverse of the temperature is not linear, indicating that the
tivities of K’ and SO:-. The newly determined log Kspfor enthalpy of reaction, AH:, varies over the temperature range
jarosite of - 11.O results in a transition pH which is 0.6 pH from 4-35°C. Therefore, a variable enthalpy and constant
units higher than calculated using the value of -9.21 from heat capacity model (Nordstrom and Munoz, 1994) was used
Kashkay et al. (1975) and Chapman et al. (1983). For a water to model the temperature dependence
194 D. Baron and C. D. Palmer

where k is the apparent rate coefficient. Integration of Eqn.


14 yields a linear equation for log (C, - C) vs. time. The
slope of the line is the apparent rate coefficient, k, and the
intercept, b, is the log of C,. The data were also plotted as
-12
(C, - C))’ vs. time and C/t vs. C to test for second-order
4 behavior but deviations from second-order behavior become
m -14 obvious in both types of plots. It appears that the dissolution
9 reaction at pH 2 and 25°C can be best described by a first-
order model with an apparent rate coefficient, k, of approxi-
-16 mately 7.9 + 0.5 X lo-’ s-l.

-16 ’ 1
6. SUMMARY
275 300 325 350 375 400
Temperature (K)
The solubility, free energy of formation, enthalpy of reac-
FIG. 7. Calculated log ion activity,product versustemperature from tion, entropy of reaction, heat capacity, and dissolution ki-
four dissolution experiments at 4, 15.25, and 35°C. initial pH 2 (W). netics of a synthetic jarosite were determined in a set of dis-
The solid line is the best fit of the variable enthalpy and constant heat solution experiments conducted in the temperature range of 4
capacity model to the experimental data extrapolated to 125°C.The
solubility product measured by Kashkay et al. ( 1975) at 100°C (0) to 35°C and pH 1.5 to 3. The rate of the dissolution reaction
is also included, but it was not used to determine the model fit. can be described by a first-order model. Equilibrium is at-
tained in the dissolution experiments after 80 to 125 days.
The solubility and free energy of formation that we deter-
mined are in excellent agreement with data from Alpers et al.
AS; - AC,,,, AH; (AC,,, log T) (1989) who studied K-Na-H,O-jarosite solid solutions. The
log&P = R - RT + R 1 (13)
Ksp reported in the present study is almost two log units lower
than the currently accepted value from Kashkay et al. (1975)
where AS: is the entropy of reaction, AC,,, is the heat capac- and Chapman et al. (1983). The higher Ksp obtained by these
ity of reaction, R is the gas constant (8.314 X lo-’ kJ mol-’ researchers may represent the solubility product of a thin sur-
K-‘), and T is the temperature in K. Fitting the temperature- face layer of hydronium-enriched jarosite that could have
dependent data to Eqn. 13 (Fig. 7) yields AH&s = -45 + S formed in their free-drift precipitation experiments. The sig-
kJ mol- ‘, AS&, = -350 ? 35 J mol-’ K-‘, and AC,,, nificantly lower solubility product obtained in this study in-
= -2.1 2 0.2 kJ mol-’ K-‘. This model gives an excellent dicates that jarosite is more stable and may occur at pH values
fit of the data (r’ = 0.999) and although only four data points higher than previously thought.
were used to determine the three parameters (AH:,,,,, The temperature dependence of the solubility product was
AS,,,, , AC,>,r), a t-test indicates that all parameters are sta- determined for the temperature range from 4 to 35°C. Based
tistically significant at the 95% confidence level. Further sup- on a solubility measurement at 100°C by Kashkay et al. (1975)
port for our results is provided by Kashkay et al. (1975) who it appears that the parameters we determined for the temper-
report a log Ksp,~,~of -15.83 for a jarosite precipitated at
100°C. Extrapolating Eqn. 13 with the parameters determined
from the experiments at temperatures between 4 and 35°C to -3.5 , I I
100°C we obtain a log Ksp,373of -15.6 that is in excellent log(C,-C)=-k t+ko -
agreement with the value measured by Kashkay et al. (1975) k = 7.9 kO.5x 10 -‘S -’
(Fig. 7). Unfortunately, we do not know the exact composition
of the jarosite precipitated at 100°C by Kashkay et al. (1975)
and if it contained a significant amount of hydronium. How-
ever, their result is consistent with our experiments and it
appears that the parameters we determined adequately de-
scribe the temperature dependence of the Ksp of jarosite over
at least the range from 4 to 100°C. -6 -
The dissolution experiment with initial pH 2 at 25°C (Fig.
4) was used to evaluate the rate of the jarosite dissolution
reaction. This experiment was sampled most frequently and
the only one with sufficient data for a reaction rate analysis. -,
1,200
0 400 600
The potassium concentrations were plotted as log (C, - C)
Time (hours)
vs. time, where C,, is the concentration at equilibrium and C
is the concentration at time t (Fig. 8). The linearity of the plot FIG. 8. Plot of the log of the equilibrium concentration of potas-
indicates that the dissolution kinetics can be described by a sium (Co) minus the concentration of potassium at time t (C) for
first-order model of the form dissolution experiment KJAR-2.0, to test for first order behavior.
Concentrations are in moUL. The solid line is the regression.The last
datapoint was not usedfor the regressionsince the difference between
dC the measuredand equilibrium concentrations is less than the analyt-
- = -k(C,, - C),
dt ical error.
Solubility of jarosite 195

ature range of our experiments can also be used to model the Keith W. J., Calk L.. and Ashley R. P. (1979) Crystals of coexisting
alunite and jarosite, Goldfield. Nevada. USGS Shorter Contribu-
solubility up to 100°C.
tions to Mineralogy and Petrology C 1-C5.
Kubisz J. (1964) A study of minerals in the alunite-jarosite group.
Polska Akad. Nauk, Prace Geol. 22, 1-93.
Acknowledgmenrs-This manuscript has benefitted from comments
Kubisz J. (1970) Studies on synthetic alkali-hydronium jarosite. I.
by Charles N. Alpers, D. Kirk Nordstrom, Eric J. Reardon, and Mark
Synthesis of jarosite and natrojarosite. Mineral. Pal. 1, 47-57.
A. Williamson. This work was supported under a grant from the U.S.
Kubisz J. (1972) Studies on synthetic alkali-hydronium jarosite. III.
Environmental Protection Agency, Office of Exploratory Research.
Infrared adsorption study. Mineral. Pal. 3,23-37.
This paper has not been reviewed by U.S. EPA and does not neces-
Kulp J. L. and Adler H. H. (1950) Thermal study of jarosite. Amer.
sarily reflect the views of the Agency.
J. Sci. 248,475-487.
Langmuir D. and Whittemore D. 0. (1971) Variations in the stability
of precipitated ferric oxyhydroxides. In Nonequilibrium Systems
Editorial handling: J. D. Rimstidt
in Natural Water Chemistry (ed. J. D. Hem); Advances in Chem-
istry Series, No. 106. pp. 209234. American Chemical Society.
Long D. T., Fegan N. E., McKee J. D., Lyons W. B., Hines M. E..
REFERENCES
and Macumber P. G. (1992) Formation of alunite, jarosite and
hydrous iron oxides in a hypersaline system: Lake Tyrell, Victoria,
Allison J. D.. Brown D. S., and Nova-Gradac K. J. (1990) MIN-
Australia. Chem. Geol. 96, 183-202.
TEQA2/PRODEFAZ, A Geochemical Assessment Model for En-
Naumov G. B., Ryzhenko I. L., and Khodakovsky I. L. (1971) Hand-
vironmental Systems: Version 3.0. U. S. Environmental Protection
book of Thermodynamic Data (translated from Russian by G. J.
Agency, Athens, GA.
Soleimami. edited by I. Barnes and V. Speltz); Washington D. C.
Alpers C. N., Nordstrom D. K., and Ball J. W. (1989) Solubility of
NTIS publication PB-226 722.
jarosite solid solutions precipitated from acid mine waters, Iron
Nordstrom D. K. (1977) Hydrogeochemical and microbiological fac-
Mountain, California, U. S. A. Sci. Geol. Bull. 42, 281-298.
tors affecting the heavy metal chemistry of an acid mine drainage
Alpers C. N., Rye R. O., Nordstrom D. K., White L. D., and King
system. Ph.D. dissertation, Stanford Univ.
B. S. (1992) Chemical, crystallographic and stable isotopic prop-
Nordstrom D. K. and Munoz J. L. (1994) Geochemical Thermody-
erties of alunite and jarosite from acid-hypersaline Australian
namics, 2nd ed. Blackwell.
lakes. Chem. Geol. 96,203-226.
Parker R. L. (1962) Isomorphous substitutions in natural and syn-
Ball J. W.. Nordstrom D. K., and Zachmann D. W. (1987)
thetic alunite. Amer. Mineral. 47, 127- 136.
WATEQ4F. A personal computer FORTRAN translation of the
Powers D. A., Rossman G. R., Schugar H. J., and Gray H. B. (1975)
geochemical model WATEQZ with revised data base. USGS Open-
Magnetic behavior and infrared spectra of jarosite. basic iron sul-
File Rept. 87-150. fate, and their chromate analogs. J. Solid State Chem. 13, I- 13.
Bladh K. W. (1982) The formation of goethite, jarosite. and alunite Ripmeester J. A., Ratcliffe C. I., Dutrizac J. E., and Jambor J. L.
during the weathering of sulfide bearing felsic rocks. Econ. Geol. (1986) Hydronium ion in the alunite-jarosite group. Canadian
77, 176- 184. Mineral. I&435-447.
Brophy G. P. and Sheridan M. F. (1965) Sulfate studies. IV. The Robie R. A. and Waldbaum D. R. (1968) Thermodynamic properties
jarosite-natrojarosite-hydronium jarosite solid solution series. of minerals and related substances at 298.15 K (25°C) and one
Amer. Mineral. 50, 112- 126. atmosphere (1 .O13 bars) pressure and at higher temperatures. U.S.
Brown J. B. (1970) A chemical study of some synthetic potassium- Geological Survey Bulletin 1259.
hydronium jarosites. Canadian Mineral. 10,696-703. Stoffregen R. E. (1993) Stability relations of jarosite and natrojarosite
Carson C. D. and Dixon J. B. (1983) Mineralogy and acidity of an at 150-250°C. Geochim. Cosmochim. Acta 57,2417-2429.
inland acid sulfate soil of Texas. Soil Sri. Sot. Amer. J. 47, 82- Stoffregen R. E. and Rye R. 0. (1992) Jarosite-water “0 and D
833. fractionations. Amer. Chem. Sot. Div. Geochem. 204, 86 (abstr.).
Chapman B. M., Jones D. R., and Jung R. F. (1983) Processes con- Sullivan P. J. and Sobek A. A. (1982) Laboratory weathering studies
trolling metal ion attenuation in acid mine drainage streams. Geo- of coal refuse. Mineral. Environ. 4, 9- 17.
chim. Cosmochim. Acta 47, 195 - 1973. Tkachenko R. I. and Zotov A. V. (1974) Ulta-acid therms of volcanic
Cox J. D., Wagman D. D., and Medvedev V. A. (1989) CODATA origin as mineralizing solutions. In Hydrothermal MineraLforming
Key Values for Thermodynamics. Hemisphere Publishing Corpo- Solutions in the Areas of Active Volcanism (ed. S. I. Naboko), pp.
ration 126- 13 1. Oxoniam Press.
Dutrizac J. E. (1983) Factors affecting alkali jarosite precipitation. Van Breemen N. (1973) Soil forming processes in acid sulfate soils.
Metall. Trans. B 14B, 531-539. In Acid Sulfare Soils; Proc. Int. Symp. on Acid Sulfate Soils (ed.
Dutrizac J. E. and Kaiman S. (1976) Synthesis and properties of H. DOST), Vol I. Publ. 18 ILRI, pp. 66- 130.
jarosite-type compounds. Canadian Mineral. 14, 15 I- 158. Vlek P. L. G., Blom T. J. M., Beek J., and Lindsay W. L. (1974)
Hartig C., Brand P., and Bohmhammel K. (1984) Fe-Al-lsomorphie Determination of the solubility product of various iron hydroxides
and Strukturwasser in Kristallen vom Jarosit-Alunit-Typ. Z. An- and jarosite by the chelation method. Soil Sci. Sot. Amer. Proc.
org. Allg. Chem. 508, 159- 164. 38,429-432.
Hladky G. and Slansky E. (1981) Stability of alunite minerals in Wagman D. D.. Evans W. H.. Parker V. B., Halow I., Bailey S. M.,
aqueous solution at normal temperature and pressure. Bull. and Schumm R. H. (1969) Selected values of chemical thermo-
Mineral. 104,468-477. dynamic properties. National Bureau of Standards Tech. Note
JCPDS (Joint Committee on Powder Diffraction Standards) (1991) 270-4. U.S. Department of Commerce.
Mineral Powder Diffraction File. International Center for Diffrac- Wilkins R. W. T., Mateen A., and West G. W. (1974) The spectro-
tion Data. Swarthmore, Pennsylvania. scopic study of oxonium in minerals, Amer. Mineral. 59,811-819.
Kashkay C. M., Borovskaya Y. B., and Badazade M. A. (1975) De- Zotov A. V., Mironova G. D., and Rusinov V. L. (1973) Determi-
termmatton of AG,?,, 0 of synthetic jarosite and its sulfate ana-
nation of AGp?,, of jarosite synthesized from a natural solution.
logues. Geochem. Intl. 12, 115- 121. Geochem. Intl. 5.577-582.

Você também pode gostar