ArticlePDF Available

Thermal expansion and decomposition of jarosite: A high-temperature neutron diffraction study

Authors:

Abstract and Figures

The structure of deuterated jarosite, KFe3(SO4)2(OD)6, was investigated using time-of-flight neutron diffraction up to its dehydroxylation temperature. Rietveld analysis reveals that with increasing temperature, its c dimension expands at a rate ~10times greater than that for a. This anisotropy of thermal expansion is due to rapid increase in the thickness of the (001) sheet of [Fe(O,OH)6] octahedra and [SO4] tetrahedra with increasing temperature. Fitting of the measured cell volumes yields a coefficient of thermal expansion, α=α0+α1 T, where α0=1.01×10−4K−1 and α1=−1.15×10−7K−2. On heating, the hydrogen bonds, O1···D–O3, through which the (001) octahedral–tetrahedral sheets are held together, become weakened, as reflected by an increase in the D···O1 distance and a concomitant decrease in the O3–D distance with increasing temperature. On further heating to 575K, jarosite starts to decompose into nanocrystalline yavapaiite and hematite (as well as water vapor), a direct result of the breaking of the hydrogen bonds that hold the jarosite structure together.
Content may be subject to copyright.
ORIGINAL PAPER
Thermal expansion and decomposition of jarosite:
a high-temperature neutron diffraction study
Hongwu Xu ÆYusheng Zhao ÆSven C. Vogel Æ
Donald D. Hickmott ÆLuke L. Daemen Æ
Monika A. Hartl
Received: 22 December 2008 / Accepted: 5 May 2009 / Published online: 24 May 2009
ÓSpringer-Verlag 2009
Abstract The structure of deuterated jarosite,
KFe
3
(SO
4
)
2
(OD)
6
, was investigated using time-of-flight
neutron diffraction up to its dehydroxylation temperature.
Rietveld analysis reveals that with increasing temperature,
its cdimension expands at a rate *10 times greater than
that for a. This anisotropy of thermal expansion is due to
rapid increase in the thickness of the (001) sheet of
[Fe(O,OH)
6
] octahedra and [SO
4
] tetrahedra with increas-
ing temperature. Fitting of the measured cell volumes yields
a coefficient of thermal expansion, a=a
0
?a
1
T, where
a
0
=1.01 910
-4
K
-1
and a
1
=-1.15 910
-7
K
-2
.On
heating, the hydrogen bonds, O1D–O3, through which the
(001) octahedral–tetrahedral sheets are held together,
become weakened, as reflected by an increase in the DO1
distance and a concomitant decrease in the O3–D distance
with increasing temperature. On further heating to 575 K,
jarosite starts to decompose into nanocrystalline yavapaiite
and hematite (as well as water vapor), a direct result of the
breaking of the hydrogen bonds that hold the jarosite
structure together.
Keywords Jarosite Neutron diffraction
Thermal expansion Decomposition Hydrogen bonds
Crystal chemistry
Introduction
Jarosite, KFe
3
(SO
4
)
2
(OH)
6
, and the related sulfates that
comprise the so-called ‘‘alunite supergroup’’ (Jambor
1999), commonly occur in acid drainage environments, as
the weathering products of sulfide ore deposits. They are
found in clays as nodules and in acid soils, where previ-
ously existing pyrite was oxidized into jarosite. They can
also precipitate from aqueous sulfate due to oxidation of
H
2
S in epithermal environments and hot springs associated
with volcanic activity (Papike et al. 2006). In 2004, jarosite
was detected by the Mars Exploration Rover (MER)
Mo
¨ssbauer spectrometer (Klingelho
¨fer et al. 2004), and it
has been interpreted as strong evidence for the occurrence
of large amounts of water (and possibly life) in the history
of Mars. A recent study using laser desorption Fourier
transform mass spectrometry revealed the presence of
organic matters (such as glycine) in several jarosite sam-
ples (Kotler et al. 2008), lending some support to the
hypothesis that life existed on Mars.
In addition to its geological importance, jarosite is of
considerable interest for its industrial applications (Dutrizac
and Jambor 2000). Specifically, in the zinc industry, Zn is
usually extracted from Zn-sulfides (such as sphalerite) by
the so-called ‘‘roast-leach-electrolysis’’ process. However,
these sulfides commonly contain Fe, typically 5–12 wt%,
which needs to be removed. Precipitation of jarosite com-
pounds has been found to be an effective means for the Fe
removal, as they form quickly and are readily filterable and
washable. This process operates at atmospheric pressure,
rather than requiring an autoclave as for many hydrothermal
processes, and is thus economical. Furthermore, the gen-
erated jarosite (in the form of mud) can be combined with
other industrial wastes such as dump ferrous slag (DFS) and
alkaline Al-surface cleaning waste (ASCW) as well as small
H. Xu (&)D. D. Hickmott
Earth and Environmental Sciences Division,
Los Alamos National Laboratory,
Los Alamos, NM 87545, USA
e-mail: hxu@lanl.gov
Y. Zhao S. C. Vogel L. L. Daemen M. A. Hartl
Los Alamos Neutron Science Center,
Los Alamos National Laboratory,
Los Alamos, NM 87545, USA
123
Phys Chem Minerals (2010) 37:73–82
DOI 10.1007/s00269-009-0311-5
portions of Portland cement or lime to produce materials for
construction applications (such as airfield runways and
levee cores) (Mymrin et al. 2005). In addition, jarosite and
its associated alunite-type phases have been proposed as
potential hosts for the long-term immobilization of radio-
active fission products and toxic heavy metals (Ballhorn
et al. 1989; Kolitsch et al. 1999).
The structure of jarosite consists of [SO
4
] tetrahedra
and distorted [Fe(O,OH)
6
] octahedra with K located in a
12-fold coordinated site (space group R
3m) (Fig. 1a)
(Menchetti and Sabelli 1976; Stoffregen et al. 2000;
Basciano and Peterson 2007). Each [Fe(O,OH)
6
] octahedron
corner-shares four hydroxyl groups with neighboring
[Fe(O,OH)
6
] octahedra and two oxygen atoms from two
[SO
4
] tetrahedra (one above the Fe and one below),
forming (001) sheets of [Fe(O,OH)
6
] and [SO
4
] perpen-
dicular to the caxis. There are two types of crystallo-
graphically distinct [SO
4
] tetrahedra: one [SO
4
] tetrahedron
pointing upward along c(c?), and the other [SO
4
] tetra-
hedron pointing downward (c-), which alternate in a
zigzag fashion along the a-axes within the (001) layer
(Fig. 1b). Each K is coordinated by 6 O atoms from [SO
4
]
tetrahedra and 6 OH groups from [Fe(O,OH)
6
] octahedra.
All 6 O atoms and all 6 OH groups are symmetrically
identical, and thus the K site has a highly symmetrical
coordination with 6 identical K–OH bonds and 6 identical
K–O bonds (Papike et al. 2006).
The unique distribution of [Fe(O,OH)
6
] octahedra
within the (001) layer coupled with the magnetic properties
of Fe
3?
makes jarosite a model compound for studying the
spin frustration in two-dimensional kagome
´lattices (com-
posed of magnetic ions located at corners of triangles that
are linked via corner-sharing) (Wills et al. 2000). Low-
temperature neutron diffraction experiments reveal that
jarosite exhibits long-range magnetic ordering when cooled
below 65 K, as evidenced by the appearance of several
magnetic reflections at hkl/2, l=odd (Inami et al. 2000).
The c-dimension of the magnetic unit cell is twice that of
the conventional unit cell, and the magnetic structure
belongs to the so-called ‘‘q =0, 120°type’’ with triangles
of the spins having only positive chirality (Inami et al.
2000). The magnetic ordering is interpreted as a result of
the coupling between the jarosite (001) layers exhibiting a
net magnetization, which is mainly due to Dzyaloshinsky-
Moriya (DM) anisotropic interactions (Grohol et al. 2003;
Yildirim and Harris 2006).
Fig. 1 a Crystal structure of
jarosite, KFe
3
(SO
4
)
2
(OH)
6
;ba
sheet of [Fe(O,OH)
6
] octahedra
and [SO
4
] tetrahedra projected
along the c-axis; cball-and-
stick representation of the
jarosite structure. Tetrahedra
represent [SO
4
] units, octahedra
represent [Fe(O,OH)
6
] units,
pink balls represent K, green
balls represent Fe, brown balls
represent S, blue balls represent
O(light blue O1 and O2;
dark blue O3), and red balls
represent H. Blue lines in aand
boutline the unit cell, and the
dash line in cmarks the
hydrogen bond between O1 and
H. In aand c, the c-axis of the
jarosite structure is vertical
74 Phys Chem Minerals (2010) 37:73–82
123
Despite the detailed structural studies of jarosite at room
and low temperatures, no information is available about its
high-temperature structural behavior. The recent discovery
of jarosite on Mars has spurred interest in its stability at
various temperatures, pressures, and aqueous conditions
(such as solution pH). A number of thermochemical studies
of jarosite and its analogues have been performed to
determine their decomposition temperatures, enthalpies of
formation, and enthalpies of dehydroxylation (Drouet and
Navrotsky 2003; Drouet et al. 2004; Forray et al. 2005;
Frost et al. 2005; Navrotsky et al. 2005). However, the
changes in the jarosite structure upon heating are still
poorly constrained. In particular, its coefficients and
mechanisms of thermal expansion remain unknown. Since
the high-temperature structural behavior of jarosite is likely
to be related to changes in its hydroxyl behavior and since
neutron scattering is sensitive to the position of hydrogen,
high-temperature neutron diffraction studies are particu-
larly useful to unravel the mechanisms of its thermal
expansion and decomposition.
In this study, we carried out in situ neutron diffraction of
jarosite using a pulsed neutron source at temperatures up to
650 K (the sample started to decompose into yavapaiite
KFe(SO
4
)
2
, hematite Fe
2
O
3
and water vapor D
2
O between
550 and 575 K). To avoid the large incoherent scattering of
neutrons by hydrogen, we synthesized deuterated jarosite,
KFe
3
(SO
4
)
2
(OD)
6
, using hydrothermal methods. Rietveld
analysis of the time-of-flight neutron data allowed deter-
mination of structural parameters as a function of temper-
ature. In particular, the atomic positions and atomic
displacement parameters of jarosite at high temperatures
have been obtained for the first time, and the structural
effects on jarosite thermal expansion and stability are
discussed.
Experimental methods
Sample synthesis
The jarosite sample used in this study was prepared via
hydrothermal methods. First, 8.1 g of Fe(NO
3
)
3
9D
2
O
(20 mmol) (Acros Organics, [99%) and 3.5 g of K
2
SO
4
(Acros Organics, [99%) were dissolved separately in
25 mL D
2
O. Second, the two solutions were mixed and
stirred thoroughly in a 100 mL Teflon cup, which was then
placed in a standard Parr autoclave. Third, the autoclave
was sealed and heated at 433 K for 3 days. After cooling
down to room temperature, the autoclave was opened, and
the contents filtered and washed with cold D
2
O. Lastly, the
resulting solid product was dried in air for one hour, placed
in a vacuum oven at 383 K overnight and then stored in a
desiccator. The product, a brown, well-crystallized powder,
was confirmed to be single-phase jarosite by powder X-ray
diffraction (Rigaku Ultima III, 40 keV, 50 mA, CuKa
radiation). The K, Fe and S contents of the sample were
measured by inductively coupled plasma atomic emission
(ICP-AE) spectroscopy. The determined weight concen-
trations are 7.66% K, 35.0% Fe and 12.6% S. The water
content, 11.7%, was determined by heating weighted
powders of the material to 723 K for *4 h and attributing
the weight loss to D
2
O. These values are very close to the
stoichiometric compositions of 7.71% K, 33.1% Fe, 12.7%
S and 11.9% D
2
O in KFe
3
(SO
4
)
2
(OD)
6
, respectively. Thus
in the following structural analysis, we treated the sample
as having the ideal formula.
Neutron diffraction
Time-of-flight neutron diffraction experiments were per-
formed at the High-Pressure Preferred Orientation (HiPPO)
beamline of the Manuel Lujan, Jr. Neutron Scattering
Center, Los Alamos National Laboratory. Sample powders
were placed in a vanadium can 0.95 cm in diameter, and
the can was mounted in an ILL-type high-temperature
furnace with vanadium heating elements and heatshields
for contamination-free diffraction data collection (Vogel
et al. 2004). Data were collected under vacuum at room
temperature and at temperatures from 350 to 650 K with an
interval of 25 K. For each temperature point, three detector
banks with nominal diffraction angles of 40°,90°and 140°
were simultaneously used. The heating rate was 5 K/min,
and the dwell time at each targeted temperature (including
an equilibration time of 5 min) was *4h.
Structure refinement
The neutron data were analyzed using the Rietveld method
with the General Structure Analysis System (GSAS) pro-
gram of Larson and Von Dreele (2000). The starting
structural parameters for KFe
3
(SO
4
)
2
(OD)
6
at 298 K were
taken from the neutron diffraction study of Menchetti and
Sabelli (1976). We then used the refined structural param-
eters at 298 K as the starting parameters for the next highest
temperature and continued this procedure systematically
with increasing temperature. For the runs at 575 and 600 K,
since a portion of the sample decomposed into yavapaiite,
hematite and water vapor, we included yavapaiite and
hematite as secondary phases in the Rietveld analyses. The
starting structural parameters for yavapaiite and hematite
were taken from the X-ray diffraction studies of Anthony
et al. (1972) and Maslen et al. (1994), respectively. For the
two highest temperature runs (625 and 650 K), only
yavapaiite and hematite were present, and thus jarosite was
excluded from the analyses. For each temperature point,
two datasets from the detectors at 2h=90°and 140°were
Phys Chem Minerals (2010) 37:73–82 75
123
simultaneously analyzed (the 40°dataset was not used
because of its relatively low resolution).
The refinements proceeded as follows: after the scale
factor and four background terms (Shifted Chebyshev
function) for each histogram had converged, lattice
parameters and phase fractions (for the runs at 575, 600,
625 and 650 K) were added and optimized. Fourteen or
eighteen additional background terms were then added for
each histogram, and the peak profiles were fitted to a TOF
profile function (Von Dreele et al. 1982). On convergence
of the preceding parameters, atomic coordinates and iso-
tropic atomic displacement parameters for K, Fe, S, O, and
D were refined, yielding R
wp
values ranging from 1.33 to
1.43%, R
p
from 0.87 to 0.98%, and v
2
from 2.8 to 3.9. The
refined unit-cell parameters, atomic coordinates, atomic
displacement parameters, and selected bond parameters are
listed in Tables 1,2,3, and 4, respectively. A representa-
tive pair of fitted patterns is plotted in Fig. 2.
Results and discussion
Stability of jarosite
Our high-temperature neutron diffraction patterns indicate
that the deuterated jarosite sample was stable up to 550 K.
However, it started to decompose into yavapaiite, hematite
and D
2
O vapor when the temperature reached 575 K:
KFe3SO4
ðÞ
2ODðÞ
6!KFe SO4
ðÞ
2þFe2O3þ3D2O:ð1Þ
As shown in Fig. 3, at 575 K, new diffraction peaks
indicative of yavapaiite and hematite appeared, and the
molar ratio for jarosite:yavapaiite:hematite obtained from
Rietveld analysis was 74.8:12.6:12.6. When the temperature
was increased to 600 K, these new peaks grew, and more
obviously, the original jarosite peaks (such as 003) became
significantly weaker. The refined molar ratio for
jarosite:yavapaiite:hematite at 600 K was 28.6:35.7:35.7.
With increasing temperature to 625 K, jarosite decomposed
completely, as revealed by the disappearance of its
diffraction peaks, and only yavapaiite and hematite were
present. Thus the onset temperature of the jarosite
dehydroxylation (T
d
) lies between 550 and 575 K. This is
generally consistent with previous thermal analyses of
potassium jarosite, which show that the mass loss due to
the dehydroxylation occurs in the temperature range 403–
603 K (Frost et al. 2005). On the other hand, as in other
hydroxyl-containing compounds such as portlandite (Xu
et al. 2007), the measured dehydroxylation temperature
can vary with sample purity, sample crystallinity and
experimental conditions such as heating rate and water
vapor pressure. These factors may account for some of the
discrepancies in the reported T
d
values for jarosite (Frost
et al. 2005; Drouet and Navrotsky 2003).
Table 1 Unit-cell parameters of deuterated jarosite and agreement
indices of the refinements
T(K) a(A
˚)c(A
˚)V(A
˚
3
)R
wp
(%) R
p
(%)
298 7.29013(6) 17.1921(2) 791.28(1) 1.33 0.87
350 7.29109(6) 17.2293(2) 793.20(1) 1.33 0.87
375 7.29275(6) 17.2514(2) 794.58(1) 1.33 0.87
400 7.29422(6) 17.2713(2) 795.82(1) 1.33 0.88
425 7.29541(7) 17.2916(2) 797.02(1) 1.33 0.89
450 7.29603(7) 17.3129(3) 798.13(1) 1.35 0.92
475 7.29644(7) 17.3316(3) 799.08(1) 1.35 0.93
500 7.29702(7) 17.3478(3) 799.96(1) 1.42 0.97
525 7.29759(7) 17.3632(3) 800.79(2) 1.43 0.98
550 7.29760(8) 17.3761(3) 801.39(2) 1.42 0.96
575 7.29775(10) 17.3854(4) 801.85(2) 1.42 0.89
Table 2 Atomic coordinates of deuterated jarosite
T(K) z(S) z(O1) x(O2) z(O2) x(O3) z(O3) x(D) z(D)
298 0.3077(2) 0.3913(1) 0.22320(7) -0.05488(5) 0.12731(7) 0.13499(6) 0.19585(8) 0.10988(5)
350 0.3070(2) 0.3907(1) 0.22323(7) -0.05518(5) 0.12741(7) 0.13500(6) 0.19555(8) 0.10978(5)
375 0.3066(2) 0.3904(1) 0.22338(7) -0.05532(5) 0.12746(7) 0.13504(6) 0.19533(8) 0.10971(5)
400 0.3061(2) 0.3901(1) 0.22339(7) -0.05547(5) 0.12751(8) 0.13497(6) 0.19523(8) 0.10966(5)
425 0.3057(2) 0.3898(1) 0.22346(7) -0.05562(5) 0.12754(8) 0.13497(7) 0.19506(8) 0.10957(6)
450 0.3053(2) 0.3895(1) 0.22347(7) -0.05573(6) 0.12755(8) 0.13502(7) 0.19490(8) 0.10959(6)
475 0.3051(2) 0.3893(1) 0.22347(7) -0.05581(6) 0.12761(8) 0.13502(7) 0.19480(9) 0.10961(6)
500 0.3045(2) 0.3889(1) 0.22349(7) -0.05581(6) 0.12765(9) 0.13491(7) 0.19467(9) 0.10967(6)
525 0.3041(3) 0.3887(1) 0.22348(8) -0.05578(6) 0.12761(9) 0.13492(8) 0.19449(9) 0.10979(6)
550 0.3038(3) 0.3885(1) 0.22355(9) -0.05562(7) 0.1276(1) 0.13487(9) 0.1942(1) 0.11000(7)
575 0.3035(4) 0.3883(2) 0.2236(1) -0.05566(9) 0.1276(1) 0.1349(1) 0.1941(2) 0.10994(9)
x(K) =y(K) =z(K) =0; x(S) =y(S) =0; x(Fe) =-y(Fe) =-z(Fe) =1/6; x(O1) =y(O1) =0; x(O2) =-y(O2); x(O3) =-y(O3);
x(D) =-y(D)
76 Phys Chem Minerals (2010) 37:73–82
123
Note that the overall intensities of the patterns at 600
and 625 K are much weaker than those at lower tempera-
tures (Fig. 3). More specifically, diffraction peaks for the
newly formed phases, yavapaiite and hematite, are broad
and not well resolved. This behavior suggests that these
phases are probably nanocrystalline in nature, presumably
due to the relatively low temperatures of their formation.
Similar behavior has been observed in simple hydroxides
such as brucite [Mg(OH)
2
], where nanocrystalline MgO
forms upon brucite dehydroxylation at 600 K (Sharma
et al. 2004).
Thermal expansion
Although jarosite has trigonal symmetry (space group
R
3m), its structure is conventionally treated in terms of a
hexagonal cell (defined by two unit-cell parameters aand c).
On heating, both aand cincrease, and thus cell volume
Valso increases (Fig. 4). However, as shown in Fig. 4a
and b (plotted on the same scale), the structural expan-
sion of jarosite occurs at a much higher rate along the
c-axis than along the a-axis and is thus highly aniso-
tropic, which is consistent with the layered nature of its
structure. To obtain the mean coefficients of thermal
expansion (CTEs), we fitted the cell-parameter data to
linear relations:
a¼7:2818 þ2:9756 105Tð2Þ
c¼16:9810 þ7:2339 104Tð3Þ
V¼779:727 þ3:9835 102Tð4Þ
The derived mean CTEs of KFe
3
(SO
4
)(OD)
6
in the
temperature range 298–575 K are: a
a
=4.0814 9
10
-6
K
-1
;a
c
=4.2066 910
-5
K
-1
; and a
V
=5.0322 9
10
-5
K
-1
. Thus the c-axis expands *10 times more
rapidly than the a-axis with increasing temperature.
The cell volume data can also be fitted to a more general
equation for thermal expansion:
VTðÞ¼V0exp ZaTðÞdT
 ð5Þ
where V
0
is the volume at a chosen reference temperature,
T
0
, and a(T) is the thermal expansion coefficient, having
the form:
aTðÞ¼a0þa1T:ð6Þ
Using T
0
=298 K, the fit yielded the following
parameters: V
0
=790.99 A
˚
3
,a
0
=1.01 910
-4
K
-1
, and
a
1
=-1.15 910
-7
K
-2
. This fit is excellent, as indi-
cated by an R
2
value of 0.995 and by the fact that the
refined V
0
is approximately the same as the measured V
0
(791.28 A
˚
3
).
Table 3 Isotropic atomic displacement parameters of deuterated
jarosite
T(K) U
iso
(K) U
iso
(S) U
iso
(Fe) U
iso
(O)
a
U
iso
(D)
298 2.9(1) 1.10(7) 0.83(2) 1.21(1) 2.72(3)
350 3.0(1) 1.19(7) 0.89(2) 1.29(1) 2.89(3)
375 3.2(1) 1.27(7) 0.94(2) 1.38(1) 3.08(3)
400 3.4(1) 1.40(7) 1.00(2) 1.46(1) 3.22(3)
425 3.5(1) 1.46(7) 1.05(2) 1.53(2) 3.38(3)
450 3.7(1) 1.54(7) 1.09(2) 1.59(2) 3.50(4)
475 3.8(1) 1.54(8) 1.12(2) 1.64(2) 3.59(4)
500 3.9(1) 1.63(8) 1.18(2) 1.67(2) 3.57(4)
525 3.8(1) 1.70(8) 1.21(2) 1.74(2) 3.60(4)
550 3.8(1) 1.77(9) 1.26(2) 1.78(2) 3.62(4)
575 3.9(2) 1.76(11) 1.34(3) 1.86(3) 3.60(6)
The unit of U
iso
:A
˚
2
/100
a
The U
iso
’s for the three O atoms are set to be equal
Table 4 Selected bond parameters of deuterated jarosite
T (K) K–O2(A
˚) K–O3(A
˚) S–O1(A
˚) S–O2(A
˚) Fe–O2(A
˚)
a
Fe–O3(A
˚)
b
D–O3(A
˚)DO1(A
˚) Fe–O3–Fe(°)
298 2.9721(8) 2.823(1) 1.437(3) 1.479(1) 2.0501(9) 1.9815(4) 0.967(1) 1.952(1) 133.77(6)
350 2.9751(8) 2.828(1) 1.442(3) 1.477(1) 2.0493(9) 1.9825(4) 0.964(1) 1.961(1) 133.69(6)
375 2.9786(8) 2.832(1) 1.447(3) 1.474(1) 2.0501(9) 1.9831(4) 0.962(1) 1.966(1) 133.67(6)
400 2.9805(8) 2.834(1) 1.450(3) 1.472(1) 2.0499(9) 1.9841(4) 0.961(1) 1.970(1) 133.58(6)
425 2.9830(9) 2.836(1) 1.454(4) 1.471(1) 2.0500(9) 1.9847(5) 0.960(1) 1.974(1) 133.55(6)
450 2.9843(9) 2.839(1) 1.459(4) 1.469(2) 2.050(1) 1.9849(5) 0.958(1) 1.979(1) 133.55(6)
475 2.9852(9) 2.842(1) 1.459(4) 1.468(2) 2.051(1) 1.9853(5) 0.957(1) 1.983(2) 133.51(7)
500 2.9860(9) 2.843(1) 1.464(4) 1.465(2) 2.053(1) 1.9863(5) 0.954(1) 1.988(2) 133.40(7)
525 2.986(1) 2.844(1) 1.469(4) 1.463(2) 2.055(1) 1.9873(5) 0.951(1) 1.993(2) 133.40(7)
550 2.986(1) 2.845(1) 1.471(5) 1.460(2) 2.059(1) 1.9866(6) 0.946(1) 2.001(2) 133.38(8)
575 2.987(1) 2.846(2) 1.474(6) 1.458(2) 2.060(2) 1.9866(8) 0.946(2) 2.003(2) 133.37(10)
a
Average of two Fe–O2 edges
b
Average of four Fe–O3 edges
Phys Chem Minerals (2010) 37:73–82 77
123
To the best of our knowledge, the obtained CTEs rep-
resent the first measurement of thermal expansion for
jarosite and its related alunite group. The a
V
value of
5.0322 910
-5
K
-1
falls within the a
V
range for common
compounds. However, it is significantly smaller than the a
V
values of many other hydroxyl-bearing minerals with a
layered structure. For example, brucite, Mg(OH)
2
, has an
a
V
of 10.9 910
-5
K
-1
(Redfern and Wood 1992), about
two times that of jarosite. On the other hand, like brucite,
jarosite exhibits large anisotropy in axial thermal expan-
sion with a much higher CTE along the caxis (normal to
the layer) than perpendicular to c. The mechanisms that
underlie this anisotropic thermal expansion are detailed
below.
Structural variation
Figure 5shows variation of isotropic displacement factors
(U
iso
) for K, Fe, S, O and D with temperature. As expected,
for a given element, its U
iso
increases with increasing
temperature. At a given temperature, U
iso
(Fe) \U
iso
(S) &
U
iso
(O) \U
iso
(D) &U
iso
(K). These trends are consistent
with the decreased bond strengths from Fe to S/O to D/K
(with their neighboring atoms), as U(= kT/f, where kis the
Boltzman constant, Tabsolute temperature, and fthe bond
force constant) is inversely proportional to the bond force
constant. Generally, the lighter the element, the weaker the
bond strength and thus the larger the U
iso
. However,
exceptions do occur, depending on the bonding configu-
ration of a given atom in the structure. In jarosite, K is
situated between the (001) [Fe(O,OH)
6
]/[SO
4
] sheets
(Fig. 1a) and thus has relatively weaker electrostatic
interactions with its adjacent O and OD. As a result, K
exhibits U
iso
values that are similarly high to those of D,
although it is much heavier.
As describe above, the jarosite structure is based on (001)
sheets of [Fe(O,OH)
6
] octahedra and [SO
4
] tetrahedra
(Fig. 1a). [Fe(O,OH)
6
] octahedra are linked via corner-
sharing, forming six- and three-membered rings perpen-
dicular to the c-axis (Fig. 1b). Each three-membered
[Fe(O,OH)
6
] ring is connected to one [SO
4
] tetrahedron
through one of the two sets of apical vertices, and
(A) 2θ= 90º
(B) 2θ= 140º
Intensity (a.u.)
1.0 4.0
3.0
2.0
d (Å)
Fig. 2 A pair of fitted neutron diffraction patterns of deuterated
jarosite collected at a2h=90°and b2h=140°at 298 K. Data are
shown as plus signs, and the solid curve is the best fit to the data. Tick
marks below the pattern show the positions of allowed reflections, and
the lower curve represents the difference between the observed and
calculated profiles
625 K: yavapaiite + hematite
600 K: jarosite + yavapaiite + hematite
575 K: jarosite + yavapaiite + hematite
Hematite 104 Yavapaiite 111/201
Jarosite 003
Intensity (arbitrary unit)
550 K: jarosite
08 282313 18 3833 4843 53 58 63.... .......
d (Å )
-
Fig. 3 Neutron diffraction patterns (2h=90°) of the deuterated
jarosite sample collected at 550, 575, 600, and 625 K. At 575 K,
jarosite started to decompose into yavapaiite and hematite, as
evidenced by the appearance of their diffraction peaks (e.g., hematite
104 and yavapaiite 111/
201). The decomposition was completed at
625 K, as indicated by the disappearance of jarosite peaks such as 003
78 Phys Chem Minerals (2010) 37:73–82
123
neighboring [SO
4
] tetrahedra point in opposite directions
(c?or c-). Therefore, among the four O atoms in a [SO
4
]
tetrahedron, three of them (O2) are each shared by one S and
one Fe, but the fourth O (O1) is bonded only to one
S (Fig. 1c). Since O1 is underbonded relative to O2, the
S–O1 distance is expected to be shorter than S–O2; structure
refinement of jarosite at room temperature shows that its
S–O1 and S–O2 distances are 1.437 and 1.479 A
˚, respec-
tively. These values fall within the observed S–O distances
for 112 refined sulfate structures, which vary from 1.394 to
1.578 A
˚(Hawthorne et al. 2000). However, the S–O2 value
of 1.479 A
˚is very close to the grand mean S–O distance,
1.473 A
˚, calculated from the 112 sulfate structures, whereas
S–O1 is significantly lower. This observation can be
explained using a formal charge model that was initially
developed to explain local structures in alkaline-earth
boroaluminates (Bunker et al. 1991). In this model, only
network-forming cations are considered. The charge dona-
ted by different cations is taken to be the cation charge
divided by the cation coordination number, as in Pauling’s
second rule (Pauling 1960) and Brown and Shannon’s
treatment of bond strengths (Brown and Shannon 1973). In
the case of jarosite, each S
6?
cation donates a charge of ?6/4
(or ?1.5), and each Fe
3?
donates a charge of ?3/6 (or ?0.5).
Because O2 is bonded to one S
6?
and one Fe
3?
, it has a net
charge of zero, resulting in a typical S–O distance associated
with O2. On the other hand, as O1 is bonded only to one S
6?
,
it receives a charge of ?1.5 and has a net charge of –0.5. To
compensate for this charge deficiency, S–O1 bond needs to
be strengthened, thereby causing the contraction of S–O1.
With increasing temperature, S–O1 increases from
1.437 A
˚at 298 K to 1.474 A
˚at 575 K (Table 4). This
behavior is consistent with the larger thermal expansion of
the c-dimension, as S–O1 is parallel to the c-axis (Fig. 1c).
On the other hand, S–O2 decreases from 1.479 A
˚at 298 K
to 1.458 A
˚at 575 K. This S–O2 shortening can be
explained in terms of [Fe(O,OH)
6
] octahedral tilting. On
heating, the Fe-O3-Fe bond angle becomes smaller (from
133.77°at 298 K to 133.37°at 575 K, Table 4). Since
each [SO
4
] tetrahedron shares three O2 atoms with a
three-membered [Fe(O,OH)
6
] ring (one O2 from each
[Fe(O,OH)
6
] octahedron) (Fig. 1b), the narrowing of the
Fe–O3–Fe angel effectively decreases the O2–O2 distance
7.36
7.40
a (Å)
7.20
7.24
7.28
7.32
A
7.16
17 34
17.38
17.42
c (
Å
)
17.22
17.26
17.30
17.
B
17.18
800
802
804
V3)
792
794
796
798
C
T (K)
280 320 360 400 440 480 520 560 600
790
280 320 360 400 440 480 520 560 600
280 320 360 400 440 480 520 560 600
Fig. 4 Variation of unit-cell parameters aa,bc, and ccell volume V
of deuterated jarosite with temperature
4.0
4.5
Å2)
3.0
3.5
K
D
Uiso (1/100
2.0
2.5
D
O
S
Fe
U
1.0
1.5
T(K)
250 300 350 400 450 500 550 600
0.5
T
Fig. 5 Variation of isotropic atomic displacement parameters (U
iso
)
of K, S, Fe, O and D in deuterated jarosite with temperature
Phys Chem Minerals (2010) 37:73–82 79
123
of the [SO
4
] tetrahedron, which causes shortening of the
S–O2 bond. Moreover, as shown in Fig. 1a, [Fe(O,OH)
6
]
octahedral layers are puckered, rather than being flat planar
(which would correspond to a Fe–O3–Fe angle of 180°).
Thus the decrease in the Fe–O3–Fe angle results in an
increase in the degree of the (001) layer puckering via
octahedral tilting. As a result, the overall structure expands
along the c-axis but contracts parallel to the (001) plane.
On the other hand, individual [Fe(O,OH)
6
] octahedra
expand with increasing temperature, as reflected by the
larger Fe–O2 and Fe–O3 distances. This causes expansion
of the structure along both a- and c-axis. It appears that the
net increase in aresulting from the thermal expansion of
[Fe(O,OH)
6
] octahedra is largely canceled by the decrease
due to the octahedral layer puckering, and thus aonly
shows a slight expansion. By contrast, both octahedral
expansion and tilting contribute to the structural expansion
along the c-axis, which, together with the S–O1 lengthen-
ing, leads to a much larger expansion along c.
Although jarosite exhibits a larger thermal expansion
along the c-axis (normal to the layer) than along a,as
observed in other hydrous minerals with a layered struc-
ture, its volume expansion coefficient (a
V
) is significantly
smaller. In other words, the overall jarosite structure is less
flexible in terms of expansion at elevated temperatures.
This behavior can be interpreted on the basis of its unique
structural characteristics. For many layered hydrous min-
erals (such as brucite [Mg(OH)
2
]), their structures can be
treated as consisting of a structural layer (e.g., the [MgO
6
]
layer in brucite) and the interlayer in which weak bonds
(e.g., van der Waals forces) are operating. Thus the thermal
expansion of these layered structures is controlled by both
the structural layer and interlayer, but the latter typically
plays a more significant role. In contrast, the jarosite
structure is comprised only of layers of [Fe(O,OH)
6
]
octahedra and [SO
4
] tetrahedra (Fig. 1a), lacking a distinct
interlayer as in other hydrous structures. The absence of the
interlayer is due to the constraint that neighboring [SO
4
]
tetrahedra linked to different [Fe(O,OH)
6
] octahedral lay-
ers must have the same height along the c-axis (there is
only one crystallographically distinct S in the unit cell of
jarosite) (Fig. 1a). Hence, the thermal expansion of jarosite
is determined solely by the flexibility of its [Fe(O,OH)
6
]/
[SO
4
] sheets, resulting in a smaller a
V
compared with those
of other layered hydrous compounds.
Despite the absence of a distinct weak-bonding inter-
layer in jarosite, the sheets of [Fe(O,OH)
6
] and [SO
4
]
polyhedra are held together by the interstitial K
?
cation via
the K–O2 and K–O3 bonds and the hydrogen bonding
between O1 and D, O1D–O3 (Fig. 1c). As in other lay-
ered hydrous structures, these bonds are weaker than the
bonds within the octahedral/tetrahedral sheets. As a result,
the K–O2, K–O3 and DO1 distances exhibit relatively
larger increases with increasing temperature (Table 4). In
particular, the DO1 attraction, which operates between a
given D (or H) and its closest O1 from the [SO
4
] tetrahe-
dron of the neighboring [Fe(O,OH)
6
]/[SO
4
] sheet (Fig. 1c),
becomes weakened, as manifested by the increase in
DO1 distance (Fig. 6a). In contrast, the O3–D bond
length shows decreases on heating (Fig. 6b), suggesting
that the O3–D bond becomes somewhat stronger. In other
words, in the O1D–O3 bonding configuration, with
increasing temperature, the O3 atoms pull the D atoms
closer, thereby effectively weakening the DO1 attraction.
Hence, the interatomic interactions of D with its neigh-
boring O atoms are interdependent and are largely driven
by the thermal motion of D at elevated temperatures.
Mechanism of jarosite decomposition
It is conceivable that the stability of jarosite is dictated by
the stability of the hydrogen bond, O1D–O3, as it, along
with K
?
, holds the structural sheets of [Fe(O,OH)
6
] octa-
hedra and [SO
4
] tetrahedra together. Once this hydrogen
(Å)
198
1.99
2.00
2.01
D...O1
194
1.95
1.96
1.97
.
A
.
0.965
0.970
B
O3-D (
Å
)
0.950
0.955
0.960
T (K)
250 300 350 400 450 500 550 600
0.940
0.945
250 300 350 400 450 500 550 600
Fig. 6 Variation of interatomic distances aDO1 and bO3–D in
deuterated jarosite as a function of temperature
80 Phys Chem Minerals (2010) 37:73–82
123
bond is broken due to high-temperature dehydroxylation,
the jarosite structure disintegrates into yavapaiite, hematite
and water vapor. More specifically, [Fe(O,OH)
6
] octahedra
become [FeO
6
] octahedra after dehydroxylation, and
one-third of the [FeO
6
] octahedra combine with [SO
4
]
tetrahedra, via corner-sharing, forming [Fe(SO
4
)
2
] sheets
parallel to the (001) plane. These sheets are linked together
by interstitial 10-coordinated K
?
, resulting in a layered
compound, yavapaiite. In the meantime, the remaining
two-thirds of the [FeO
6
] octahedra are connected via edge-
sharing to form gibbsite-type octahedral layers, and the
latter are stacked, via face-sharing, along the c-axis,
forming hematite. Given the structural relations among
jarosite, yavapaiite and hematite, there may be certain
topotactic relations between the decomposed jarosite and
newly formed yavapaiite and hematite. Specifically, the
layered nature of all three phases may result in the fol-
lowing relation: (001)
jarosite
//(001)
yavapaiite
//(001)
hematite
or
c
jarosite
//c
yavapaiite
//c
hematite
. This type of topotactic reaction
(i.e., structurally controlled) mechanism has been found
responsible for the thermal decomposition of many min-
erals including hydroxides, oxyhydroxides and carbonates
(Sharma et al. 2004; Floquet and Niepce 1978). The
occurrence of topotactic relations among jarosite, yavap-
aiite and hematite, however, would require verification by
other techniques such as high-resolution transmission
electron microscopy.
Conclusions
We have studied the stability and structural behavior of
deuterated jarosite in the temperature range 298–650 K
using neutron diffraction in conjunction with Rietveld
analysis. Our results show that jarosite is stable up to
550 K, above which it starts to decompose into nanocrys-
talline yavapaiite and hematite. With increasing tempera-
ture, both the aand cdimension of jarosite expand, but the
latter expands at a rate *10 times larger, as is consistent
with the layered nature of its structure. On the other hand,
because of the lack of a distinct weak-bonding interlayer
between adjacent (001) sheets of [Fe(O,OH)
6
] octahedra
and [SO
4
] tetrahedra, the volume expansion coefficient of
jarosite is significantly smaller than those of many other
hydroxyl-bearing minerals with a layered structure. At a
given temperature, the amplitudes of thermal vibration of
D and K are much larger than those for Fe, O and S,
implying their weaker bonding with surrounding atoms.
Correspondingly, on heating, the DO1 distance of the
hydrogen bond O1D–O3 increases, which suggests
weakened hydrogen bonding between neighboring (001)
tetrahedral/octahedral sheets. By contrast, the O3–D bond
becomes stronger with increasing temperature, a trend also
observed in simple hydroxides such as portlandite (Xu
et al. 2007).
Acknowledgments We thank P.J. Heaney and an anonymous
reviewer for helpful comments, and M.S. Rearick for carrying out
compositional analysis of the jarosite sample. This work has benefited
from the use of the Lujan Neutron Scattering Center at LANSCE,
which is funded by the Department of Energy’s Office of Basic
Energy Sciences. Los Alamos National Laboratory is operated by Los
Alamos National Security, LLC, under DOE Contract DE-AC52-
06NA25396.
References
Anthony JW, McLean WJ, Laughon RB (1972) The crystal structure
of yavapaiite: a discussion. Am Mineral 57:1546–1549
Ballhorn R, Brunner H, Schwab RG (1989) Artificial compounds of
the crandallite type: a new material for separation and immo-
bilization of fission products. Sci Basis Nucl Waste Manage
12:249–252
Basciano LC, Peterson RC (2007) Jarosite-hydronium jarosite solid-
solution series with full iron site occupancy: mineralogy and
crystal chemistry. Am Mineral 92:1464–1473. doi:10.2138/am.
2007.2432
Brown ID, Shannon RD (1973) Empirical bond-strength–bond-length
curves for oxides. Acta Crystallogr A 29:266–282. doi:10.1107/
S0567739473000689
Bunker BC, Kirkpatrick RJ, Brow RK (1991) Local structure of
alkaline-earth boroaluminate crystals and glasses: I. Crystal
chemical concepts—structural predictions and comparisons to
known crystal structures. J Am Ceram Soc 74:1425–1429. doi:
10.1111/j.1151-2916.1991.tb04123.x
Drouet C, Navrotsky A (2003) Synthesis, characterization and
thermochemistry of K–Na–H
3
O jarosites. Geochim Cosmochi-
mica 67:2063–2076. doi:10.1016/S0016-7037(02)01299-1
Drouet C, Pass KL, Baron D, Draucker S, Navrotsky A (2004)
Thermochemistry of jarosite-alunite and natrojarosite-natroalu-
nite solid solutions. Geochim Cosmochimica 68:2197–2205. doi:
10.1016/j.gca.2003.12.001
Dutrizac JE, Jambor JL (2000) Jarosite and their application in
hydrometallurgy. In: Alpers CN, Jambor JL (eds) Sulfate
minerals: crystallography, geochemistry, and environmental
significance, reviews in mineralogy and geochemistry, vol 40.
Mineralogical Society of America, Washington DC, pp 405–452
Floquet N, Niepce JC (1978) Three-fold transformation twin in the
topotactic decomposition of cadmium carbonate crystals. J Mater
Sci 13:766–776. doi:10.1007/BF00570511
Forray FL, Drouett C, Navrotsky A (2005) Thermochemistry of
yavapaiite KFe(SO
4
)
2
: formation and decomposition. Geochim
Cosmochimica 69:2133–2140. doi:10.1016/j.gca.2004.10.018
Frost RL, Weier ML, Martens W (2005) Thermal decomposition of
jarosites of potassium, sodium and lead. J Therm Anal Calorim
82:115–118. doi:10.1007/s10973-005-0850-z
Grohol D, Nocera DG, Papoutsakis D (2003) Magnetism of pure iron
jarosites. Phys Rev B 67:064401-1-13
Hawthorne FC, Krivovichev SV, Burns PC (2000) The crystal
chemistry of sulfate minerals. In: Alpers CN, Jambor JL (eds)
Sulfate minerals: crystallography, geochemistry, and environ-
mental significance, reviews in mineralogy and geochemistry,
vol 40. Mineralogical Society of America, Washington DC, pp
1–112
Inami T, Nishiyama M, Meegawa S, Oka Y (2000) Magnetic structure
of the kagome
´lattice antiferromagnet potassium jarosite
Phys Chem Minerals (2010) 37:73–82 81
123
KFe
3
(OH)
6
(SO
4
)
2
. Phys Rev B 61:12181–12186. doi:10.1103/
PhysRevB.61.12181
Jambor JL (1999) Nomenclature of the alunite supergroup. Can
Mineral 37:1323–1341
Klingelho
¨fer G, Morris RV et al (2004) Jarosite and hematite at
Meridiani Planum from Opportunity’s Mo
¨ssbauer spectrometer.
Science 306:1740–1745. doi:10.1126/science.1104653
Kolitsch U, Tiekink ERT, Slade PG, Taylor MR, Pring A (1999)
Hinsdalite and plumbogummite, their atomic arrangements and
disordered lead sites. Eur J Mineral 11:513–520
Kotler JM, Hinman NW, Yan B, Stoner DL, Scott JR (2008) Glycine
identification in natural jarosites using laser desorption Fourier
transform mass spectrometry: implications for the search for life
on mars. Astrobiology 8:253–266. doi:10.1089/ast.2006.0102
Larson AC, Von Dreele RB (2000) GSAS—general structure analysis
system. Los Alamos National Laboratory Report No. LAUR 86–
748. Los Alamos National Laboratory, Los Alamos
Maslen EN, Strel’tsov VA, Strel’tsova NR, Ishizawa N (1994) Syn-
chrotron X-ray study of the electron density in alpha Fe
2
O
3
. Acta
Crystallogr B 50:435–441. doi:10.1107/S0108768194002284
Menchetti S, Sabelli C (1976) Crystal chemistry of the alunite series:
crystal structure refinement of alunite and synthetic jarosite.
N Jahrb Min Monatsh H 9:406–417
Mymrin VA, Ponte HA, Impinnisi PR (2005) Potential application of
acid jarosite wastes as the main component of construction
materials. Construct Build Mater 19:141–146. doi:10.1016/j.
conbuildmat.2004.05.009
Navrotsky A, Forray FL, Drouet C (2005) Jarosite stability on Mars.
Icarus 176:250–253. doi:10.1016/j.icarus.2005.02.003
Papike JJ, Karner JM, Shearer CK (2006) Comparative planetary
mineralogy: Implications of martian and terrestrial jarosite. A
crystal chemical perspective. Geochim Cosmochim Acta 70:
1309–1321. doi:10.1016/j.gca.2005.11.004
Pauling L (1960) The nature of chemical bond, 3rd edn. Cornell
University Press, Ithaca
Redfern SAT, Wood BJ (1992) Thermal expansion of brucite,
Mg(OH)
2
. Am Mineral 77:1129–1132
Sharma R, Mckelvy MJ, Bearat H, Chizmeshya AVG, Carpenter RW
(2004) In situ nanoscale observations of the Mg(OH)
2
dehydr-
oxylation and rehydroxylation mechanisms. Philos Mag
84:2711–2729. doi:10.1080/14786430410001671458
Stoffregen RE, Alpers CN, Jambor JL (2000) Alunite-jarosite
crystallography, thermodynamics, and geochronology. In: Alpers
CN, Jambor JL (eds) Sulfate minerals: crystallography, geo-
chemistry, and environmental significance, reviews in mineral-
ogy and geochemistry, vol 40. Mineralogical Society of
America, Washington DC, pp 453–479
Vogel SC, Hartig C, Lutterotti L, Von Dreele RB, Wenk HR,
Williams DJ (2004) Texture measurements using the new
neutron diffractometer HIPPO and their analysis using the
Rietveld method. Powder Diffr 19:65–68. doi:10.1154/1.
1649961
Von Dreele RB, Jorgensen JD, Windsor CG (1982) Rietveld
refinement with spallation neutron powder diffraction data.
J Appl Cryst 15:581–589. doi:10.1107/S0021889882012722
Wills AS, Harrison A, Ritter C, Smith RI (2000) Magnetic properties
of pure and diamagnetically doped jarosites: model kagome0
antiferromagnets with variable coverage of the magnetic lattice.
Phys Rev B 61:6156–6169. doi:10.1103/PhysRevB.61.6156
Xu H, Zhao Y, Vogel SC, Daemen LL, Hickmott DD (2007)
Anisotropic thermal expansion and hydrogen bonding behavior
of portlandite: a high-temperature neutron diffraction study.
J Solid State Chem 180:1519–1525. doi:10.1016/j.jssc.2007.03.004
Yildirim T, Harris AB (2006) Magnetic structure and spin waves in
the Kagome jarosite compound KFe
3
(SO
4
)
2
(OH)
6
. Phys Rev B
73:214446. doi:10.1103/PhysRevB.73.214446
82 Phys Chem Minerals (2010) 37:73–82
123
... The single-crystal hexagonal diffraction patterns remain stable until the temperature reaches 420°C, when ring patterns of yavapaiite begin to form. This phase transformation is attributed to dehydroxylation of jarosite (Xu et al., 2010). By 570°C, the diffraction patterns for both samples show a combination of ring-patterns (suggesting nanocrystalline materials) and single crystal diffraction patterns difficult to completely index. ...
... Our in-situ TEM results reveal that both hypogene and supergene jarosites undergo phase transformation at 420°C under vacuum, likely associated with dehydroxylation, compatible with previous thermogravimetric analysis (TGA) in air (Alpers et al., 1989;Vasconcelos, 1999b;Stoffregen et al., 2000;Drouet et al., 2003Drouet et al., , 2004Frost et al., 2005;Desborough et al., 2010;Waltenberg, 2012;Chen, 2018), but are not consistent with results of thermal studies on synthetic jarosite by Xu et al. (2010). The latter authors interpret neutron diffraction patterns from a synthetic jarosite sample to indicate that during vacuum heating jarosite is stable up to 550 K (277°C), above which it starts to decompose into nanocrystalline yavapaiite and hematite. ...
Article
Jarosite [KFe3(SO4)2(OH)6] occurs both as a hydrothermal mineral or as the product of weathering and chemical sedimentation. It has been used in ⁴⁰Ar/³⁹Ar geochronology to date water-rock interaction and weathering processes on the surfaces of Earth and Mars, but the lack of information about Ar diffusivity parameters relevant to specific types of jarosites makes the interpretation of geochronological results tentative. We have filled this gap by investigating Ar diffusion parameters in representative supergene and hypogene jarosites. Detailed diffusion studies were carried out on a hypogene jarosite sample from Gilbert, Nevada, and two supergene jarosite samples from Baiyin, China. The diffusion studies were accompanied by in-situ heating investigations in a transmission electron microscope to directly determine the thermal stability and the phase transformations that jarosite undergoes under progressive heating under ultra-high vacuum. The TEM results suggest that jarosite is stable under vacuum up to ∼ 400 °C, when it undergoes phase transition to yavapaiite [KFe(SO4)2] and hematite (Fe2O3). Incremental-heating experiments reveal average diffusion parameters of Ea = 138.6 ± 4.2 kJ/mol and ln(Do/a²) = 9.9 ± 0.9 ln(s⁻¹) for hypogene jarosite; Ea = 110.3 ± 3.2 kJ/mol and ln(Do/a²) = 5.7 ± 0.7 ln(s⁻¹) for one supergene jarosites; and Ea = 141.2 ± 7.9 kJ/mol and ln(Do/a²) = 11.3 ± 1.7 ln(s⁻¹) for the other supergene jarosite sample from the same weathering profile. Jarosite closure temperatures depend strongly on sieve size. For samples between 500-200 µm (grain size usually used for samples in Ar geochronology), at a cooling rate 100 °C·Ma⁻¹, the closure temperatures are 143 ± 18 °C for hypogene and 105 ± 8 and 113 ± 14 °C, respectively, for the two supergene jarosites. Forward modelling of incremental-heating results predicts that coarse-grained hypogene jarosite is retentive of Ar below 50 °C for 100 Ma and below 25 °C for 4 Ga. Densely packed supergene jarosite grains larger than 200 µm are suitable for ⁴⁰Ar/³⁹Ar geochronology at the timescales suitable for investigating water-rock interaction at the surface of Earth and Mars. Fine-grained, porous jarosites require detailed diffusion analyses prior to geochronology due to possible high Ar losses over Ma timescales. The absence of jarosites older than ∼40 Ma on Earth suggests that jarosite may require continuous exposure to acid oxidizing conditions, and that it does not survive burial and exhumation. Therefore, the occurrence of jarosite on Earth and Mars may identify segments of the planets’ surfaces continuously exposure to acid-oxidizing conditions since jarosite precipitation.
... It can also be successfully synthesized by the reaction of KOH with FeOOH in H 2 SO 4 solution, and the thermal decomposition of yavapaiite begins at approximately 500°C (Ferenc et al., 2005). Another route to obtain yavapaiite is the dihydroxylation of jarosite (Chio et al., 2010;Xu et al., 2010), which is a sulfate mineral that is usually found in acid soils or in clays as nodules, from the oxidation of pyrite (Vithana et al., 2014;Xu et al., 2010). ...
... It can also be successfully synthesized by the reaction of KOH with FeOOH in H 2 SO 4 solution, and the thermal decomposition of yavapaiite begins at approximately 500°C (Ferenc et al., 2005). Another route to obtain yavapaiite is the dihydroxylation of jarosite (Chio et al., 2010;Xu et al., 2010), which is a sulfate mineral that is usually found in acid soils or in clays as nodules, from the oxidation of pyrite (Vithana et al., 2014;Xu et al., 2010). ...
Article
During oxygen pressure acid leaching to release vanadium from black shale, a large amount of Fe is also leached, which is detrimental to the following purification and enrichment processes. To suppressed the Fe leaching, this study attempts to clarify the precise conditions for yavapaiite (KFe(SO4)2) formation and apply the abovementioned strategy to oxygen pressure acid leaching process of black shale. When the acid concentration was 10–20 vol%, a maximum Fe precipitation efficiency of approximately 81% could be achieved in the K2SO4-Fe2(SO4)3-H2SO4 solution system under the following conditions: temperature, 190 °C or higher; Fe³⁺ concentration, 0.40–0.60 mol/L; potassium to iron molar ratio, 2.0–2.5; time, 4–5 h. X-ray diffraction and scanning electron microscopy-energy dispersive spectroscopy analyses showed that yavapaiite could be formed when the H2SO4 concentration and temperature reached 10 vol% and 150 °C, respectively. Adjusting the K⁺ concentration in the leaching solution was demonstrated to an effective method to generate yavapaiite during the oxygen pressure acid leaching for black shale. The migration of Fe can be inferred as pyrite to Fe³⁺ to krausite/yavapaiite and finally to yavapaiite.
... However, one can create a one dimensional form of expansivity or compressibility to derive either linear thermal expansivity or linear compressibility (Xu et al. 2009b(Xu et al. , 2009a. Most zircontype structure compounds will respond to the external variables of temperature or pressure anisotropically in that the rate of expansion or contraction of the a-and c-axis are not equivalent. ...
Article
Full-text available
Zircon structure-types are ternary oxides with an ideal chemical formula of ATO4 (I41/amd), where usually A are lanthanides and actinides, with T = As, P, Si, V. This structure accommodates a diverse chemistry on both A- and T-sites, giving rise to more than seventeen mineral end-members of five different mineral groups, and forty-five synthetic end-members. Because of their diverse chemical and physical properties, the zircon structure-types are of interest to a wide variety of fields and may be used as ceramic nuclear waste forms and aeronautical environmental barrier coating, to name a couple. To support advancement of their applications, many studies have been dedicated to the understanding of their structural and thermodynamic properties. The emphasis in this review will be on recent advances in the structural and thermodynamic studies of zircon structure-type ceramics, including pure endmembers (i.e., zircon (ZrSiO4), xenotime (YPO4)), and solid solutions (i.e., ErxTh1-x(PO4)x(SiO4)1-x). Specifically, we offer an overview on the crystal structure, its variations and transformations in response to non-ambient stimuli (temperature, pressure and radiation), and correlation to thermophysical and thermochemical properties. We hope this review will promote further applications of these ceramic materials.
... The primary aim of this study was to examine structural transformation and evolution when pyrite undergoes decomposition or, more precisely, desulfurization upon heating under vacuum using in situ high-temperature neutron diffraction [10,11,12,13,14]. In particular, the composition and structure of the nonstoichiometric pyrrhotite (Fe 1Àx S) vary systematically with increasing temperature and/or time. ...
Article
To study thermal desulfurization of pyrite (FeS2), we conducted in situ neutron diffraction experiments in the temperature range 298–1073 K. On heating, pyrite remained stable up to 773 K, at which it started to decompose into pyrrhotite (Fe1−xS) and S2 gas. Rietveld analysis of the neutron data from 298 to 773 K allowed determination of the thermal expansion coefficient of pyrite (space group Pa\(\bar 3\)) to be αV = 3.7456 × 10−5 K−1, which largely results from the expansion of the Fe–S bond. With further increase in temperature to 1073 K, all the pyrite transformed to pyrrhotite (Fe1−xS) at 873 K. Unit-cell parameters of Fe1−xS (space group P63/mmc) increase on heating and decrease on cooling. However, the rates in cell expansion are larger than those in contraction. This hysteresis behavior can be attributed to continuous desulfurization of pyrrhotite (i.e., x in Fe1−xS decreases) with increasing temperature until the stoichiometric troilite (FeS) was formed at 1073 K. On cooling, troilite underwent a magnetic transition to an orthorhombic structure (space group Pnma) between 473 and 573 K. In addition, using differential thermal analysis (DTA) and thermogravimetric analysis (TGA) implemented with a differential scanning calorimeter, we performed kinetic measurements of pyrite decomposition. Detailed peak profile and Arrhenius (k = A exp(−Ea/RT)) analyses yielded an activation energy Ea of 302.3 ± 28.6 kJ/mol (based on DTA data) or 302.5 ± 26.4 kJ/mol (based on TGA data) and a ln(A) of 35.3 ± 0.1.
Article
The passivation layer was proven to be an inevitable factor that prevents chalcopyrite leaching. Sulfur and jarosite are the commonly investigated substances in passivation layer research. In this study, the distribution characteristics of leaching solution in the gap between chalcopyrite and jarosite and between chalcopyrite and sulfur were investigated deeply at a molecular level. And the equilibrium distribution configurations were calculated. When polyvinyl pyrrolidone (PVP) was added, a significant layer formed near chalcopyrite. It prevented the accumulation of SO4²⁻ and thus prevent the formation of the passivation layer near chalcopyrite. Additionally, the increased diffusion coefficient of H2O molecules indicated the increased permeability and mobility of the leaching solution between chalcopyrite and passivation layer by PVP. Moreover, the increased diffusion coefficients of SO4²⁻ between chalcopyrite and sulfur referred to the enhanced dissolution of sulfur. Thus, chalcopyrite leaching was enhanced by PVP. The results provide a new view for the deep understanding of the enhancement mechanism of surfactants in chalcopyrite hydrometallurgy.
Article
Full-text available
The crystal structure of moganite from the Mogán formation on Gran Canaria has been re-investigated using high-resolution synchrotron X-ray diffraction (XRD) and X-ray/neutron pair distribution function (PDF) analyses. Our study for the first time reports the anisotropic atomic displacement parameters (ADPs) of a natural moganite. Rietveld analysis of synchrotron XRD data determined the crystal structure of moganite with the space group I2/a. The refined unit-cell parameters are a = 8.7363(8), b = 4.8688(5), c = 10.7203(9) Å, and β = 90.212(4)°. The ADPs of Si and O in moganite were obtained from X-ray and neutron PDF analyses. The shapes and orientations of the anisotropic ellipsoids determined from X-ray and neutron measurements are similar. The anisotropic ellipsoids for O extend along planes perpendicular to the Si-Si axis of corner-sharing SiO4 tetrahedra, suggesting precession-like movement. Neutron PDF result confirms the occurrence of OH over some of the tetrahedral sites. We postulate that moganite nanomineral is stable with respect to quartz in hypersaline water. The ADPs of moganite show a similar trend as those of quartz determined by single-crystal XRD. In short, the combined methods can provide high-quality structural parameters of moganite nanomineral, including its ADPs and extra OH position at the surface. This approach can be used as an alternative means for solving the structures of crystals that are not large enough for single-crystal XRD measurements, such as fine-grained and nanocrystalline minerals formed in various geological environments.
Article
Although continued growth in unconventional oil and gas production is generally projected, its long-term growth potential and sustainability have significant uncertainties. A critical problem is the low hydrocarbon recovery rates from shale and other tight formations using the horizontal well drilling and hydraulic fracturing techniques: < 10% for tight oil and ~ 20% for shale gas. Moreover, the production rate for a given well typically declines rapidly within one year. The low recoveries and declining production of shale oil and gas reservoirs are apparently related to the small porosity (a few to a few hundred nm) and low permeability (10⁻¹⁶–10⁻²⁰ m²) of shale matrix, which make the enclosed hydrocarbon fluids difficult to access. Hence, to enhance the hydrocarbon recovery from shale matrix, it is essential to study its nanopore structure and confined fluid behavior. Small- and ultra-small-angle neutron scattering (SANS and USANS) have emerged as a powerful method for characterizing shale nanopore structure and confined fluid behavior. Owing to neutrons' high penetrating ability and high sensitivity to hydrogen (and its isotope, deuterium), SANS/USANS can probe inside shale samples to characterize nanopores from 1 nm to 10 μm in size and be readily combined with sample environmental cells to examine the fluid (hydrocarbon and water – frack fluid) behavior at relevant pressure-temperature (P-T) conditions. In this review article, an introduction is first given on the characteristics of shale matrix and the uniqueness of SANS/USANS compared with conventional methods. Then current studies on shale nanopore structure and confined fluid properties using SANS/USANS are summarized. Finally, an outlook and perspective on future research in this emerging area will be offered.
Article
This study investigated the sulfidation and sulfur fixation of jarosite residues during reduction roasting in the presence of carbon. The effects of roasting temperature and carbon dosage were investigated based on thermodynamic calculation. The results indicated that more than 98 pct of zinc contained in the residue was converted into zinc sulfides, and more than 91 pct of sulfur was fixed in the roasted residue. Carbon addition promoted not only the sulfidation of zinc but also the fixation of sulfur, thereby eliminating SO2 emission. The growth of sulfide particles was strongly influenced by roasting temperature. The size of sulfide particles significantly increased when the temperature was above 1173 K (900 °C) because of the formation of a liquid phase during the roasting process. However, high temperature could increase the consumption of carbon powder.
Article
A combination of time-of-flight neutron diffraction and synchrotron X-ray powder diffraction has been used to investigate the thermal expansion of a synthetic deuterated natrojarosite from 80 to 440 K under ambient-pressure conditions. The variation in unit-cell volume for monoclinic jarosite over this temperature range can be well represented by an Einstein expression of the form V = 515.308 (5) + 8.5 (4)/{exp[319 (4)/ T ] − 1}. Analysis of the behaviour of the polyhedra and hydrogen-bond network suggests that the strength of the hydrogen bonds connected to the sulfate tetrahedra is instrumental in determining the expansion of the structure, which manifests primarily in the c -axis direction.
Article
Full-text available
Na–H3O jarosite was synthesized hydrothermally at 413 K for 8 days and investigated using single-crystal X-ray diffraction (XRD) and electron microprobe analysis (EMPA). The chemical composition of the studied crystal is [Na0.57(3) (H3O)0.36 (H2O)0.07]A Fe2.93(3) (SO4)2 (OH)5.70 (H2O)0.30, and Fe deficiency was confirmed by both EMPA and XRD analysis. The single-crystal XRD data were collected at 298 and 102 K, and crystal structures were refined in space group \( R\overline{3}m\). The room-temperature data match structural trends of the jarosite group, which vary linearly with the c axis. The low-temperature structure at 102 K shows an anisotropic decrease in the unit cell parameters, with c and a decreasing by 0.45 and 0.03 %, respectively. Structural changes are mainly confined to the A site environment. Only minor changes occur in FeO6 and SO4 polyhedra. The structure responds upon cooling by increasing bond length distortion and by decreasing quadratic elongation of the large AO12 polyhedra. The structural parameters at low temperature follow very similar patterns to structural changes that correspond to compositional variation in the jarosite group, which is characterised by the flexibility of AO12 polyhedra and rigidity of Fe(OH)4O2–SO4 layers. The most flexible areas in the jarosite structure are localized at AO12 edges that are not shared with neighbouring FeO6 octahedra. Importantly, for the application of XRD in planetary settings, the temperature-related changes in jarosite can mimic compositional change.
Article
Full-text available
The lattice parameters of brucite, Mg(OH)2, have been measured between 20 and 310-degrees-C at atmospheric pressure. The thermal expansion is strongly anisotropic with a high coefficient of linear expansion (alpha3) parallel to the z axis. This leads to an overall volume coefficient of expansion (average of 10.9 x 10(-5)/K) that is several times greater than estimates based on amphiboles and other hydrous minerals. Our results for the individual expansion coefficients are a alpha1 = alpha2 = 1.73 x 10(-5) - 0.13 x 10(-7) T +/- 0.17 x 10(-5), alpha3 = 3.23 x 10(-5) + 2.65 x 10(-7) T +/- 0.32 x 10(-5), and alpha(vol) = 6.70 x 10(-5) + 2.37 x 10(-7) T +/-0.67 x 10(-5), where T is the temperature between 20 and 310-degrees-C.
Article
Structure factors for synthetic haematite, alpha-Fe2O3, have been measured for two small crystals using focused lambda = 0.7 angstrom synchrotron radiation. The structure factors from the two data sets are consistent. Approximate symmetry in the concordant densities, related more closely to the Fe-Fe geometry than to the nearest-neighbour Fe-O interactions, is similar to that in the corundum alpha-Al2O3 Structure. Deformation density maxima are located at the midpoint of the Fe-Fe vector along the c axis, on a common face for 0-octahedra, perpendicular to c. Maxima also occur at the midpoint of the Fe-Fe vector bisecting the edges of the 0-octahedra. These results are in accordance with theoretical predictions for metal-metal bonding. Space group R3cBAR, hexagonal, M(r) = 159.7, a = 5.0355 (5), c = 13.7471 (7) angstrom, V = 301.88 (7) angstrom3, Z = 6, D(x) = 5.270 MgM-3, mu0.7 = 13.68 mm-1, F(000) = 456, T = 293K, R = 0.019, wR = 0.021, S = 5.55 for 368 unique reflections in the most accurate data set 2.
Article
Minerals of the crandallite group (Ca, Sr. Ba, Pb, REE) H Al 3 ((po 4 ) 2 /(OH) 6 ) exhibit manifold possibilities for element substitutions. They occur as hydrothermal precipitates together with tin tungsten mineralization [1], as placer minerals [2], neo-formed minerals in lateritic profiles [3, 4] and lake sediments [5]. An easy and cheap method of synthesizing an artificial crandallite-type material was developed. By adding 4.2 g of crandallite to an aqueous solution containing 300 g/l sodium nitrate, 0.6 g/l europium nitrate and 0.6 g/l thorium nitrate, europium and thorium were effectively separated from the solution. This was also demonstrated for Ce. La, Nd, Gd, Zr, Ba and Sr. Artificial crandallite-type material laden with strontium was subjected to a Soxlet test. A dissolution of strontium of 0.06 to 0.08 wt.% with reference to the total strontium content resulted. As shown by Herold [6] the substitution of Ca by larger bivalent or higher charged ions results in an increase in the thermodynamic and thermal stability of artificial crandallites.
Article
The crystal structures of hinsdalite, PbAl3[(P0.69,S0.31)O4)]2(OH,H2O)6, and of a plumbogummite containing a small amount of arsenic, PbAl3[(P0.95,As0.05)O4)]2(OH,H2O)6, have been refined in space group R3m, with a = 7.029(4) and c = 16.789(4) Å, and a = 7.039(5) and c = 16.761(3) Å, respectively. The refinements, using 258 (plumbogummite: 297) observed reflections with I≥ 3σ (1), led to R =3.0 % (3.7 %) and RW = 3.0 % (3.2 %). Both minerals have the beudantite/crandallite structure type with hinsdalite being characterised by disordered (P,S)O4 tetrahedra with an average (P,S)-O distance of 1.52 Å. The Pb atoms in both minerals are displaced from the origin and are disordered about their sites, as in other Pb containing members of the alunite family (beudantite, kintoreite, and plumbojarosite). The disorder of the Pb atoms is confined to the (0001) plane; in hindsdalite, Pb is at (0.0312, 0.0312, 0.0), while in plumbogummite it is at (0.0409, 0.0409, 0.0). Pb-O distances average 2.79 Å in both minerals. The hydrogen-bonding networks are modelled with the help of bond-valence summations.
Article
The alunite supergroup consists of more than 40 minerals with the general formula DG 3( T O4)2(OH,H2O)6, wherein D represents cations with a coordination number greater or equal to 9, and G and T represent sites with octahedral and tetrahedral coordination, respectively (Smith et al. 1998). The supergroup is commonly subdivided into various groups, but the simplest primary subdivision is on the basis of the G cations. For all of the minerals in the supergroup, the dominant G cation is trivalent; most of the minerals have G represented by Fe3+ or Al3+, but exceptions are the rare minerals gallobeudantite, in which G is Ga3+, and springcreekite, in which G is V3+ (Table 1⇓). Thus, the primary grouping adopted here is on whether formula Fe3+ exceeds or is subordinate to Al3+. The hierarchical sequence in mineralogy seems to be variable, but here the decreasing sequence is given as supergroup, family, group, and subgroup. Minerals with Fe3+ > Al3+ are referred to as belonging to the jarosite family, and those with A13+ > Fe3+ are allocated to the alunite family. View this table: Table 1. Minerals of the alunite supergroup. Subdivision of the alunite and jarosite families has also been variable; Scott (1987), for example, used seven groups, Novak et al. (1994) used six, Gaines et al. (1997) used four, and Mandarino (1999) used three. The arbitrary decision here is to use three groups, which differ from those of Mandarino (1999) but which, in general, indicate whether sulfate, phosphate, or arsenate predominates in the T O4 tetrahedra. The three groups are the alunite group, in which T O4 is dominated by SO4, the crandallite group, in which (PO4) is …
Article
Stoichiometrically pure jarosites of the formula AFe3(OH)6(SO4)2 with A=Na+, K+, Rb+, and NH+4 have been afforded by a newly developed redox-based, hydrothermal method. The jarosites exhibit an intralayer antiferromagnetic exchange interaction (-829 K<ΘCW<-812 K) and transition temperatures for long-range order (LRO) (61 K<TN<65 K) that are essentially insensitive to the size of the A+ ion. A cusp at TN in the ac susceptibility curve is frequency independent. The origin of LRO is consistent with coupling of jarosite layers exhibiting a net magnetization, which arises from an anisotropy developed, most likely, from the Dzyaloshinsky-Moriya (DM) interaction. A canted intralayer spin structure, which is a consequence of the DM interaction, is signified by a remanent magnetization (˜53 K<TD<˜58 K), the magnitude of which depends on crystallite size. X-ray single crystal analyses of the pure Fe3+ jarosite compounds reveal that the kagomé layers are structurally invariant with those of their Cr3+ and V3+ relatives. This structural homology allows the sign and magnitude of exchange coupling within kagomé layers to be correlated to the different orbital parentages engendered by the M3+ d-electron count. Infrared studies show the presence of H2O within the kagomé layers of alkali metal and hydronium ion Fe3+ jarosites prepared by conventional precipitation methods; conversely, H2O is absent within the kagomé layers of jarosites prepared by the new redox-based hydrothermal methods. These results suggest that the absence of LRO in (H3O)Fe3(OH)6(SO4)2 is due to structural and magnetic disorder arising from proton transfer from the interlayer hydronium ion to the bridging hydroxide ions of the kagomé layers.