You are on page 1of 21

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/223254911

Formation of alunite, jarosite and hydrous iron oxides in a hypersaline


system: Lake Tyrrell, Victoria, Australia

Article  in  Chemical Geology · March 1992


DOI: 10.1016/0009-2541(92)90128-R

CITATIONS READS

75 537

6 authors, including:

David Long Mark E Hines


Michigan State University University of Massachusetts Lowell
119 PUBLICATIONS   3,415 CITATIONS    215 PUBLICATIONS   4,532 CITATIONS   

SEE PROFILE SEE PROFILE

Phillip Macumber
University of Melbourne
42 PUBLICATIONS   1,180 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Al Zubarah: World Heritage Site, Qatar (archaeology, history) View project

dsrB genes for sulfate reducing bacteria View project

All content following this page was uploaded by Phillip Macumber on 02 December 2018.

The user has requested enhancement of the downloaded file.


Chemical Geology, 96 (1992) 183-202 183
Elsevier Science Publishers B.V., Amsterdam

Formation of alunite, jarosite and hydrous iron oxides in a


hypersaline system: Lake Tyrrell, Victoria, Australia

D.T. Long a'~, N.E. Fegan a'~, J.D. McKee a, W.B. Lyons b'a, M.E. Hines b and P.G. Macumber c
~Geological Sciences, Michigan State University, East Lansing, M148824, USA
blnstitute for Earth, Oceans. and Space, University of New Hampshire, Durham, NH 03824-3589, USA
CDepartment of Water Resources, Melbourne, Vic. 3000, Australia
(Accepted for publication August 30, 1991 )

ABSTRACT

Long, D.T., Fegan, N.E., McKee, J.D., Lyons, W.B_, Hines, M.E. and Macumber, P.G., 1992. Formation ofalunite, jarosite
and hydrous iron oxides in a hypersaline system: Lake Tyrrell, Victoria, Australia. In: W.B. Lyons, D.T. Long, A.L.
Herczeg amd M.E. Hines (Guest-Editors), The Geochemistry of Acid Groundwater Systems. Chem. Geol., 96:183-
202.

Alunite (KAI3(SO4)2(OH)6) and jarosite (KFe3(SO4)2(OH)6) are common weathering products ofaluminosilicates
and pyrite. Long-range transport of the constituents forming these minerals and the subsequent deposition of alunite and
jarosite in an evaporite setting have not been previously documented. Such conditions for the occurrence of alunite and
jarosite were investigated in a hypersaline system where acid groundwater enriched in K, AI, Fe (11I) and SO4 enter a salt-
playa lake. Sediment cores were studied by thin section, XRD and SEM-EDS. Groundwater and pore-water chemistries
were also analyzed. The results show that alunite and jarosite occur together or as individuallayers and cements within the
top 20 cm of the sediments, where the groundwater is most concentrated due to evaporation. Hydrous Fe-oxides also occur
as cements or vein fillings with the alunile and jarosite, but are distributed throughout the 70-cm sediment column studied.
The results are consistent with a model in which alunite and jarosite precipitate as a result of the evaporation of water.
Thermodynamic modeling indicates that the pore water in the playa sediments maybe in equilibrium with alunite, jarosite
and hydrous Fe-oxides.

1. Introduction and van Everdingen, 1987). Alunite forms


during both hypogene and supergene pro-
Alunite ( K A I 3 ( S O 4 ) e ( O H ) 6 ) and jarosite cesses, whereas jarosite forms mainly during
( K F e 3 ( 5 0 4 ) 2 ( O H ) 6 ) f o r m u n d e r high Eh, l o w the latter (Brown, 1971; Hladky and S!ansky,
p H conditions and are often associated with 1981; Stoffregen and Alpers, 1987; Rye et al.,
acid-mine drainage, weathering of sulfide ore 1989). Hypogene jarosite has been found to
deposits, hydrothermal alteration, oxidation of occur at several hydrothermal deposits, but is
pyrite, or weathering in acid soils (Hemley et less common than hypogene alunite in zones
al., 1969; van Breemen, 1973; Knight, 1977; of acid-sulfate hydrothermal alteration (C. AI-
Nordstrom, 1982; Sullivan et al., 1986; Michel pers, pers. commun., 1989). In low-tempera-
ture environments the formation of these min-
~AI1 correspondence should be directed to senior author. erals has also been related to microbial
Present addresses: oxidation of iron and sulfur (Adams and Ha-
~Department of Geological Sciences, California State
University-Hayward, Hayward, CA 94542, USA.
jek, 1978; Ivarson et al., 1979). The mecha-
aHydrology/Hydrogeology Program, University of Ne- nisms controlling the formation of alunite are
vada, Mackay School of Mines, Reno, NV 89557, USA. of interest because of its potential for seques-

0009-2541/92/$05.00 © 1992 Elsevier Science Publishers B.V. All rights reserved.


184 D,T. LONG ET AL.

tering A1 and SO4 in soils and its importance formation of alunite and jarosite (Goldbery,
as a source of Al for industry (Adams and Ra- 1978, 1980).
wajfih, 1977; Adams and Hajek, 1978; Ivarson The principal mechanism for the formation
et al., 1979; Dutrizac and Jambor, 1987). Both ofjarosite and alunite is thought to be by dis-
minerals have been studied as indicators of the solution/re-precipitation reactions involving
occurrence and potential mass of sulfide de- pyrite and aluminosilicates. In some cases these
posits (Bladh, 1982) and for their ability to minerals precipitate after "local" migration
scavenge elements from solution (Ivarson et (on the order of I m) of the constituents (i.e.
al., 1979; Dutrizac and Jambor, 1987 ). K, A1, Fe, SO4) (Hemley et al., 1969). Long-
Alunite and jarosite comprise a family of distance migration of the constituents prior to
minerals with the general formula their precipitation as alunite or jarosite min-
AB3(XO4)E(OH)6, where A is a large cation erals has not been previously documented, with
(e.g., K +, Na +, Rb +, NH4 +, Ag +, H30 +, the exception of Michel and van Everdingen
Ca 2+, Pb 2+, Ba 2+, Sr 2+, Ce 3+ ), B is an octa- ( 1987 ), who have demonstrated such a case for
hedrally coordinated cation (e.g., AIs+, Fe 3÷, a jarosite deposit in shale. In this case jarosite
Cu 2+, Zn 2+) and XO~4- is an anion (e.g., is precipitating from a groundwater seep. The
SO42-, PO43-, AsO 3-, CrO 2- ). Thus, the for- constituent elements originate from water-rock
mation of these minerals can affect the geo- reactions along the groundwater flow path.
chemistry of many elements. The classifica- In the southern half of the Australian conti-
tion of the minerals in the family is based on nent, acid groundwater occurs over large areas
whether AI 3+ or Fe 3+ is the dominant cation and migrates long distances (Macumber,
in the B site, forming the alunite and jarosite 1983 ). This regional groundwater carries large
supergroups, respectively (Scott, 1987). Ex- amounts of Fe, Si and A1 (Macumber, 1983;
tensive solid solution between and within these Mann, 1983 ). In areas where this groundwater
supergroups has been demonstrated experi- discharges into playa lakes, evaporite minerals
mentally and by direct measurement of natu- form as the water evaporates (Long et al., 1992
ral minerals (Brophy et al., 1962; Hartig et al., in this issue ). Alunite, jarosite and hydrous Fe-
1984; Scott, 1987). oxides are frequently found with the evaporite
The source of acidity in solutions that pre- minerals. The association of alunite and jaros-
cipitate alunite or jarosite can be the oxidation ite with hypersaline environments in which
of sulfide, particularly pyrite or H2S, precipi- evaporite minerals are precipitating has been
tation of hydrous Fe-oxides, and possibly large little studied (Lock, 1988).
inputs of acid rain into poorly buffered sys- This study investigates the occurrence of al-
tems (Sass et al., 1965; Hemley et al., 1969; unite and jarosite in hypersaline sediments of
Nordstrom, 1982; Mann, 1983). The oxida- Lake Tyrrell, Australia. An evaporative origin
tion of pyrite also supplies Fe 3÷ and SO4. Ad- for these minerals is proposed. Their associa-
ditional sources of SO4 can be acid-sulfate soil tion with Fe-oxides and the possibility of so-
water, acid rain and dissolution of sulfate min- lution-mineral equilibrium in the alunite-ja-
erals (van Breemen, 1973). The acid weath- rosite and alunite-j arosite-Fe-oxide
ering of clays and other aluminosilicate min- assemblages are explored.
erals mobilizes A1 for alunite formation as well
as K, which is a constituent of both alunite and 2. Study area
jarosite. K also may be derived from cation-
exchange reactions. Both SOn and K can origi- The geologic history, sedimentology and hy-
nate from seawater and their concentration in drogeochemistry of the Lake Tyrrell system has
solution by evaporation could be related to the been discussed by Teller et al. (1982), Mac-
ALUNITE,JAROSITEANDHYDROUSIRONOXIDESIN A HYPERSALINESYSTEM 185

umber (1983, 1992 in this issue), Bowler aquifer, but this trend is not related to the
(1986), Bowler and Teller (1986), and Long grain-size differences of the sand.
et al. (1992 in this issue); and will be only Within the study area the Parilla is a con-
briefly discussed here. Lake Tyrrell is a 26 × 7 fined to semi-confined aquifer. It is mostly
km salt-playa lake located in the Murray Basin overlain by the Plio-Pleistocence lacustrine
of NW Victoria, Australia (Fig. 1 ). Lake Tyr- Blanchetown Clay and underlain by the Upper
rell became dry ~ 32 kyr BP, but is covered Oligocene-Lower Miocene marine Geera Clay.
with 1O's of cm of water for ~ 3 months of the The Blanchetown Clay, which is light to olive
year. The lake deposits are up to 5 m thick and grey with scattered sand lenses, is thickest (20
consist of clay, silt and sand covered by an m) in the northern area of Lake Tyrrell and is
ephemeral halite-gypsum crust (30-60 cm). absent towards the south. In the southern area
Decaying algal mats occur locally on the lake the Tyrrell beds directly overlie the Parilla
floor. Sand. The Geera Clay is dark grey to black,
The lake beds are underlain by a large aqui- ~ 100 m thick and less permeable than the
fer system formed by the Parilla Sand. This Blanchetown Clay in the Lake Tyrrell region.
aquifer is Late Miocene to Pliocene in age. In Two smaller playa lakes, Wahpool and Tim-
the study area the aquifer averages 60 m thick. boram, occur east of Tyrrell. The regional
The quartz sand appears to have been depos- groundwater flow in the Parilla Sand is west-
ited in a marginal marine environment and ward (Fig. 1 ), interacting with the three playa
contains heavy minerals and organic matter. lakes as springs or as stream input. Lake Tim-
The upper half of the Parilla is coarse-grained boram is at the highest elevation and is an open
sediment, while the sediment of the bottom or flow-through lake, whereas Lakes Tyrrell and
half is more fine grained. Most of the organic Wahpool are closed lakes. Groundwater enters
matter and pyrite occur toward the base of the Lake Tyrrell from springs around the lake

B7 88 Groundwoter Flow Line

S rlng •

:o,o\ " S cT-9 I "-- f2 ,AKE


,A ',...... :_ ~ ~ t'~ J - - t (. ") TIMBORAM

54 B2 groundwater ,N, I ~ J
B2 / ', ~g' • ',~-"~
~o26 6032
• . .e2 5S-B
-.} ) ~ \
j

I, 6026
\ .....
..
J~/ Wells ~ / - " Tyrrell Creek
o s Jo ,"- K) _.---.J ':
B 1 I I ~ w n n f ' ~ TCST e

Groundwater Flow Line

Fig. 1. Study area location showing well distribution, general path of regional groundwater flow and zone of non-acid
groundwater.
186 D.T. LONG ET AL.

shore. A major groundwater divide occurs ~ 10 line) to sample the low-pH, oxic regional
km west of Lake Tyrrell. Thus, groundwater groundwater entering the lake as springs on the
entering Lake Tyrrell along the western shore western shore (Fig. 2). These piezometers are
has its origin in the regional groundwater flow l m deep and cross the spring zone. Two of the
from the east (Fig. 1 ). piezometers are outside the spring zone and
The groundwater in the Tyrrell system has sample the Lake Tyrrell reflux brine (Fig. 2).
been characterized by salinity and location as In order to characterize the spring zone water
three major types (Macumber, 1983): further, a second piezometer line (NPZ line)
( 1 ) Regional groundwater from the Parilla was established across the spring zone ~ 150
Sand which is similar in salinity to seawater m north of the first line (Fig. 2 ). The chemical
(3.5%). The water is oxic with pH < 4 and data indicate that the Lake Tyrrell reflux brine
enters Lake Tyrrell as seeps along the western was not sampled by this new piezometer line.
shore (Fig. 1 ). Pore water in the spring zone was also col-
(2) Brine which saturates the Parilla Sand lected utilizing in situ pore-water samplers
aquifer below Lake T~crrell and is formed by called "sippers" (Fig. 2). These are basically
reflux of evapo-concentrated water. This water lysimeters made of Teflon ® with a porous Te-
is more saline (25%) than regional groundwa- flon ® sleeve (2-cm diameter) which can be
ter. The water is suboxic to anoxic with near- placed at various depths in the sediment
neutral pH. (Hines et al., 1992 in this issue). A vacuum
(3) Brine beneath Lakes Tyrrell and Wah- using an inert gas (N2) was drawn on the "sip-
pool which is also formed by evapo-concentra- per" and pore water was collected in situ. Sip-
tion of regional water. The salinity of this brine pers were placed at each site in 0-3-, 4-9-, 10-
(12%) is about half that of the Tyrrell reflux 15- and 15-20-cm depth intervals. The data set
brine. This water is also suboxic to anoxic with includes 89 samples from wells, piezometers
near-neutral pH and enters Lake Tyrrell from and sippers.
springs along the eastern shore (Fig. 1 ). The wells and the piezometers were sampled
Long et al. ( 1992 in this issue) conclude that with either a Teflon ® bailer or a peristaltic
these three types of water masses have evolved pump. The piezometers were pumped dry be-
from a common parent with sea salt as the fore sampling. Once retrieved, all samples for
source for the solutes. They suggest that pro-
cesses that occur early in the evolution of the
water such as calcrete and gypcrete formation
affect the chemical signature of the water. Aci-
dification of the water appears to happen early
in the geochemical evolution of the water and
is initiated by the oxidation of pyrite. Locally,
acid conditions are enhanced by the precipita-
tion of alunite, jarosite and hydrous Fe-oxides,
and evaporation.

3. Methods

The wells established by Macumber ( 1983 ) o.. ,,,oL,.r. ,.o


were used to sample the three types of ground-
water discussed and are shown on Fig. 1. Mac- Fig. 2. Spring zone with locations of piezometers a n d
umber ( 1983 ) put in a line of piezometers (TP sippers.
ALUNITE, JAROSITE AND HYDROUS IRON OXIDES IN A HYPERSALINE SYSTEM 187

A1, Fe and SO4 analyses were maintained and graphite furnace (AASG) using a 1:100 dilu-
processed under an N2 atmosphere in glove tion. Standards for AI analysis were prepared
bags. All samples were immediately filtered with Fe and C1 added in similar concentra-
through an acid-rinsed Teflon ® filtering sys- tions to the sample being analyzed. It is as-
tem using acid-rinsed Nuclepore ® 0.4-/~m fil- sumed that the dissolved AI obtained by this
ters for trace-metal and minor-element sam- technique is mainly monomeric. This was in
ples and Millipore ® 0.45-/1m filters for the part confirmed by comparing the results from
major elements. Except for the samples for al- the AASG technique to those obtained using
kalinity, SO4 and other anion determinations, the A1 extraction technique of Barnes ( 1975 ).
all samples were acidified to a pH < 2 using This latter technique was used to selectively
ULTREX ® HNO3. Samples were stored in extract monomeric AI from test samples. The
acid-rinsed polyethylene bottles. These bottles results show that similar A1 concentrations
were rinsed thoroughly with deionized-dis- were detected by both techniques.
tilled water before use. Samples for AI deter- To examine sediment mineralogy, 60-cm co-
minations were handled in a similar manner to res were taken from the spring zone along each
the methods described by Mackin and Aller piezometer line using 9-cm-diameter polyvi-
(1984). These samples were acidified in the nylchloride (PVC) pipe. Sediment was ex-
laboratory 24 hr prior to A1 analysis in order to truded from the core and subsampled in the
minimize potential contamination from the field. Sediment mineralogy and specifically al-
sample bottle and to maximize recovery of unite, jar®site and hydrous Fe-oxides were
monomeric A1 (J.E. Mackin, pers. commun,, identified by X-ray diffraction (XRD) (Ri-
1987). The ULTREX ® HNO3 for acidifica- gaku ® Geigerflex D/Max-Ia) on bulk samples
tion of the AI samples had to be cleaned by sub- using Cu-K, radiation. Alunite, jar®site and the
boiling distillation in a Teflon ® evaporation Fe-oxides were also examined visually and
system. Lyons et al. (1992 in this issue) fur- chemically by scanning electron microscopy
ther describe the clean procedures used in this (SEM) using energy-dispersive spectrometry
study. (EDS) (JEOL ® JSM-T20 with Tracor North-
Chloride was determined by Mohr titration ern ® EDS). During the sectioning of the sedi-
and bromide was determined by colorimetry ment cores, selected sediments were dried and
using a modification of the chloramine-T vacuum impregnated with EPO-KWICK ®
method proposed by Presley (1971) as de- epoxy cement for preparation of thin sections.
scribed by Long and Gudramovics (1982). The chemical modeling program PHREEQE
Other analytical methods and results for the (Parkhurst et al., 1980) was used to study the
major- and minor-element determinations are potential for thermodynamic equilibrium
decribed by Long et al. (1992 in this issue). among alunite, jar®site and hydrous Fe-oxides
The major constituents of alunite, jar®site and in the system. PHREEQEis a mass transfer/spe-
the hydrous Fe-oxides in the water samples ciation code that uses the data base from
were analyzed as follows. Total Fe and Fe (II) Truesdell and Jones (1974) and Ball et al.
were determined by the Ferrozine ® technique ( 1987 ) which has its origin in the Garrels and
(St®®key, 1970; Gibbs, 1976; Lyons et al., Thompson (1962) model for the determina-
1989) on filtered samples immediately after tion of activities of solutes in aqueous solu-
collection. Sulfate was determined turbidi- tions. We recognize that the ability of this
metrically (Hines et al., 1992 in this issue) and computer code to define activities at high ionic
K by atomic absorption spectrometry (AAS) strengths is debatable and that the Pitzer-based
using a 1:210 dilution with a NaCI-LaCI3-HCI (e.g., Pitzer, 1979) codes such as PHRQPITZ
matrix added. A1 was analyzed by AAS with (Plummer et al., 1988) are more suitable for
188 D.T LONGET AL.

the study of high ionic strength solutions the sediment. Few feldspars were actually
(Harvie and Weare, 1980). However, Pitzer- identified in the sediments from the spring
based codes are presently limited in their abil- zone. Heavy minerals identified in the sedi-
ity to describe metals such as Fe 3+ and AI 3+, ments include tourmaline, zircon, epidote and
particularly in those systems involving oxida- dark opaques probably including magnetite, il-
tion-reduction processes (Reardon and Be- menite, hematite and others. The clays identi-
ckie, 1987) and were deemed not suitable for fied are illite and kaolinite (Long et al., 1992
this study. in this issue). Clay/silt layers in the lower sec-
tion of cores contain numerous fractures at
4. Results and discussion various angles to bedding. These fractures are
stained red and yellow and appear to have been
4. I. Occurrence of alunite, jarosite and conduits for the upward flow of the oxidized
hydrous iron oxides groundwater from the Parilla Sand. This ob-
servation is consistent with the hydrologic
In addition to the sediment in the spring zone model proposed for the spring zone system by
along the western shore (Fig. 1 ), sediment was Macumber ( 1983, 1992 in this issue ).
also examined at the southern end of Lake Most of the red, purple and yellow streaks
Tyrrell where Tyrrell Creek enters the lake and and stains in the sediments are hydrous Fe-ox-
at spring zones along the eastern shore. Of these ides that occur either as discrete layers or len-
three areas, only in the western spring zone did ses within the sediment or as intergranular ce-
alunite, jarosite and hydrous Fe-oxides occur ments. The yellow cement was identified by
together. The sediments in the spring zone XRD as goethite. This was the only Fe-oxide
consist of bedded silts, sands, and some clays conclusively identified by XRD; although mi-
overlain by a halite-gypsum crust. The crust is nor amounts oflepidocrocite and hematite may
thickest towards the center of the lake outside have been present. The red and purple streaks
the spring zone where it is > 40 cm thick at were X-ray amorphous. Fig. 3 is a photomicro-
piezometer TP0 (Fig. 2). Below the salt crust, graph using crossed nicols of goethite cement
or top 2 cm of the sediment where the salt is surrounding quartz grains and a microcline
absent, there are bands of reduced organic grain. The microcline appears corroded, but the
matter ~ 5 cm thick. timing of the corrosion is not known.
Within the spring zone, the near-surface sed- Like the Fe-oxides, alunite and jarosite oc-
iment occurs as thin layers (0. 1-5 cm) of silt cur as discrete layers within the sediment and
and some clay colored different shades of yel- as cements. The alunite is white to greyish
low, white and grey. Below these sediments are white while the jarosite is a brilliant yellow.
layers of sand and clay, varying in thickness Both minerals were frequently misidentified in
between 10 and 50 cm, which are generally grey the field as gypsum and Fe-oxide, respectively.
with patches and streaks of red, yellow and When the XRD data of the alunite and jarosite
purple. The sands are compositionally mature, are compared to standard XRD data, these
containing generally in excess of 90% quartz. minerals appear to be more like the K-rich end-
In thin section, the quartz is mostly monocrys- members than the Na-rich end-members (A1-
talline and nonundulose. The quartz grains pers et al., 1989, 1992 in this issue). There-
tend to be angular rather than rounded, so their fore, they would have the formulas of
texture is not mature. In addition, quartz grains KA13(SO4)2(OH)6 and K _ F : e 3 ( S O a ) 2 ( O H ) 6 ,
coated with Fe-oxides are disseminated respectively. The Na-enriched forms of these
throughout the sediments. These grains give the minerals were not detected in the sediments.
appearance of feldspar in hand specimens of The predominance of K over Na as the A cat-
ALUNITE, JAROSITE AND HYDROUS IRON OXIDES IN A HYPERSALINE SYSTEM 189

Fig. 3. Photomicrographunder crossed nicolsofgoethite cement surrounding a microclinegrain (scale bar= 0.1 ram).

ion was qualitatively confirmed by EDS scans larized light ofjarosite cementing quartz grains.
(presented below). Alpers et al. ( 1992 in this Periods of infilling of the jarosite cement are
issue) provide data on unit cell dimensions of seen as layering. Under crossed polarizers the
alunite and jarosite based on powder X R D jarosite cement appears to be homogeneous.
which further confirm the dominance of the K- No evidence of direct replacement of either
rich end-member. pyrite or aluminosilicates by jarosite was seen
Alunite and jarosite tend to occur in the top in thin section. It was not possible to obtain a
20 cm of the sediment. This is in contrast to thin section of the alunite suitable for photo-
the Fe-oxides, which occur at all depths within microscopy. The alunite cement was not strong
the 70-cm length of the cores. Individual layers enough to hold the sediment together during
of alunite and jarosite cements, usually < 5 cm, thin section preparation and did not allow the
are as thin as 0.3 cm. Neither these layers nor epoxy cement to penetrate the sediment enough
the depth of occurrence of the alunite and ja- to bond the sediment and alunite. In addition,
rosite cements is laterally continuous through- part of the alunite cement disaggregated dur-
out the spring zone. ing this preparation. X R D data showed the
Although the regional groundwater entering presence of halite in the alunite cement, which
the spring zone are oxic (Macumber, 1983, may have caused the apparent dissolution.
1992 in this issue; Long et al., 1992 in this is- Under a binocular microscope, alunite, ja-
sue), the sediments through which the ground- rosite and the Fe-oxides were clearly seen to
water flows in the spring zone do not appear to cement the quartz grains of the sediment. Both
have been completely oxidized. In addition to alunite and jarosite cements encircled quartz
the reduced sediment layer at or near the sur- grains coated with Fe-oxides. More iron-coated
face, grey unoxidized sediment occurs quartz grains were found within the jarosite
throughout the cores in layers from 0.3 to 7 cm cement than in the alunite cement. This obser-
thick. Frequently, jarosite and alunite occur vation is consistent with the observation that
immediately below these unoxidized sediment hydrous Fe-oxides t~:nd to occur more fre-
layers. quently with the jarosite than with the alunite
Fig. 4 is a photomicrograph using plane-po- in the sediments. Water was clearly seen to dis-
90 D.T. LONGET AL.

Fig. 4. Photomicrograph under plane-polarized light ofjarosite cement (scale bar= 0. I m m ) .

Fig. 5_ SEM photomicrograph of alunite (scale bar= 1/~m).

aggregate the alunite cement, but not the jaros- 4.2. Characterization of alunite, jarosite and
ite cement. This is consistent with the ob- hydrous iron oxides
served association of halite with the alunite.
However, we cannot make the statement that Figs. 5-8 show typical SEM images of alu-
all alunite in the spring zone is associated with nite, jarosite and the hydrous Fe-oxides. Alu-
halite whereas all jarosite is not, because we nite has a sugary texture composed of fine-
have XRD data showing all three minerals in grained, relatively equant crystals of < 1 /lm
one sample. cross-section (Fig. 5). Jarosite, on the other
ALUNITE, JAROSITE AND HYDROUS IRON OXIDES IN A HYPERSALINE SYSTEM 191

Fig. 6. SEM photomicrograph ofjarosite (scale bar= 2 #m).

Fig. 7. SEM photomicrograph of X-ray amorphous Fe-oxide (scale bar= 1/zm).

hand, tends to occur as hexagonal platelets ides which have precipitated on quartz. Fig. 8
which are > 1/zm in cross-section and --, 0.2/zm is a SEM image of halite. The EDS scans of the
thick (Fig. 6). The X-ray amorphous Fe-ox- flakes of material on the halite showed the
ides (Fig. 7) occur as individual spheroids ~ 1 flakes to be composed ofAl, K and S (Fig. 9).
/zm in diameter which also coalesce into more This indicates that the flakes are alunite and is
massive aggregates. This figure shows Fe-ox- consistent with the observation o f the close as-
192 D.T. LONG ET AL.

Fig. 8. SEM photomicrographof halite with alunite flakes (scale bar= 2 ltm).

peaks of Fe. It is not clear if these small peaks


[A)
AI K
~ Fa
reflect solid solution or mechanical mixing of
s L Jk jarosite and alunite. This uncertainty has also
(8) j~ been noted by Alpers et al. (1992 in this is-
j~-.Z sue). In many cases, both alunite and jarosite
were detected by X R D in samples that ap-
(c} I peared monomineralic by macroscopic inspec-

[D}.~L ~~c, tion. This might suggest that the jarosite and
alunite mainly occur as a mechanical mixture
rather than as a solid solution. However, more
work on the stoichiometry of these minerals
needs to be done before this problem can be
completely resolved. The color of jarosite-al-
Energy(keV) unite mixtures is discussed by Alpers et al.
Fig. 9. EDS scans of minerals shown in Figs. 5-8: (a) al- ( 1992 in this issue).
unite: (b) jarosite; (c) iron oxide; and (d) alunite flakes
on halite. 4.3. Aqueous solution chemistry

sociation of alunite and halite from the thin Table 1 summarizes the aqueous solution
section and X R D results. chemistry in regional and spring zone ground-
The EDS results (Fig. 9) clearly show that water for the major constituents of alunite, ja-
alunite and jarosite are composed of the K end- rosite and hydrous Fe-oxides including pH; Br
member. This was found to be the case for the is included because in provides a conservative
EDS scans of all the alunite and jarosite sam- natural tracer. The regional groundwater in-
pies. Small peaks of Al were found in thejaros- cludes only samples with pH < 5.5, the ap-
ire scans, and the alunite scans contained small proximate division between the acid and near-
r-

>
z

©
z
TABLE 1 ©
x
Concentrations (in mg 1-' ) of selected constituents

Al Fe(III) K Br SO4 pH

RGW SZW RGW SZW RGW SZW RGW SZW RGW SZW RGW SZW

X 16.3 68.2 8.91 3.74 279 769 151 460 5,442 19,838 3.99 3.47
M 2.4 41.0 5.43 1.94 240 691 110 398 3,084 19,035 3.75 3.56
Mo 2.4 10.2 2.77 1.62 240 1,050 70.1 362 4,150 3,285 3.70 3.60
GM 2.96 38.4 2.77 1.62 220 639 115 373 3,662 14,674 3.93 3.44
R 0.05-76.6 3.19-247 0.1-26.7 0.05-20.5 11.6-738 190-2,080 3.37-467 107-1,190 154-18,789 2,027-66,112 2.9-5.38 2.6-4.25
CF 0.05-7.2 1.6-18.3

R G W = regional groundwater, n = 35; S Z W = spring zone ground- and pore waters, n = 23; X = mean; M = median; Mo = mode; GM = geometric mean;
R = range; CF = concentration factor relative to seawater.
194 D.T. LONG ET AL.

5.7
j ~ . . d i , stituents. Fe(IlI) concentrations are approxi-
mately the same in both the regional and spring
5.2 SQGWQtQr zone groundwater, as are the pH-values.
@v Q p - c u r ve
4.7 A plot of Cl vs. Br (Fig. 10) shows that the
4.2 solute chemistry of the Lake Tyrrell ground-
U * wells water is similar to evaporated seawater (Long
3.7 s!ppers
0
• p~ezos et al., 1992 in this issue). Therefore, Br can be
3.2 used as an indicator of the degree to which a
2.7 n n n I water sample has been concentrated from a
0.5 1.25 2 2.75 3.5 seawater-like source. Concentration factors
Log Br ( m g / 1 ) (CF) are calculated as (Br in a sample ) / (Br
in seawater). These concentration factors are
Fig. 10. Iog,o CI vs. log,o Br. Seawater evaporation curve
from MacCaffrey et al. (1988). listed in Table 1 along with the Br data.
Plots of the major constituents vs. Br are
neutral pH groundwater masses. The spring shown in Fig. 11. Similar to C1, SO4 shows a
zone chemistry combines the sipper and pi- trend typical for evaporated seawater (Fig.
ezometer data within this zone (Fig. 2). Ex- 11 a). Chloride and sulfate geochemisty in the
cept for Fe 3+, the highest concentrations of the Lake Tyrrell system has been shown to be con-
constituents are found in the spring zone trolled by halite and gypsum precipitation, re-
groundwater where evaporation is occurring. spectively (Long et al., 1992 in this issue).
This indicates that evapo-concentration is in Potassium (Fig. 1 lb) also covaries with
part controlling the concentration of the con- bromide, but plots below the evaporative trend

5.0 1.6
e y a p - c u r ve . ~ • •
Q•
4.0 0.6 oS °

0.1
, ells
ob~ zx s!ppers • D
3.0 ~ -0.4
lag • plezos o
o ~ -0.9
* A C
2.0 , , , , i , , o . . . . . . , . . . . ,
-1.4 •
4.0 2.6 . . . . . . . . . . . . . . . . .
S@O WO t Qr"

3.0
Q v£Tp - c u r " v e
1.6

V* •
..
• Q-oo_
.

0.6 •° • ol o•

0
2.0 ~,* * wells
• ~ s!ppers -0.4 qbo • • •
• puezos
B D
1.0 *i i I I -1.4 L . . . . , . . •.
. . • • . I . . . . t

1.5 2 2.5 3 3.5 i .5 2 2.5 3 3.5


Log B r ( m g / 1 ) Log B r ( r a g / l )

Fig. 11. logno constituent vs. iog,o Br: (a) SO4; (b) K; (c) FeS+; and (d) AI. Seawater evaporation curves from Mac-
Caffrey et al. (1988).
ALUNITE, JAROSITE AND H Y D R O U S IRON OXIDES IN A HYPERSALINE SYSTEM 195

3.0 be expected to be minor (Jones et al., 1977).


2.0
Mineral precipitation could also account for
• ,
the K removal. Graphical analysis and chemi-
1.0 cal modeling data indicate that these waters are
• • dDqb •
0.0
Qo • not saturated with salts of K typical of evapo-
• 1110 00
• • • rative environments (Long et al., 1992 in this
-1.0
A
issue). Therefore, it is suggested that the pre-
-2.0 i
I .
.
.
.
. .
. . i
t .
.
.
.
.
.
.
.
]
i
.
. .
.
.
.
.
. . .
1
.
,
.
1 .
.
. 1 .
I
¸
. . . .
.
.
w
. .
. .
.
.
.
.
.
w
cipitation of alunite and jarosite may in part
2.0
account for the removal of K from solution
tag 1.5
during evaporation (Long et al., 1992 in this
1.0 Ooo •
• • issue).
0.5
O • 0 ..'.. The geochemistry of A1 and Fe is affected not
0.0
• gO only by evaporation and mineral precipita-
-0.5
tion, but also by pH and, in the case of Fe, Eh
-1.0
• B controls. Consequently, a simple one-to-one
-1.5
2 3 4 5 6 7 8
relationship between the logio concentration of
these elements and lOgl0 Br concentrations
pH
would not be expected. This is the case for Fe 3+
Fig. 12. Iogzo constituent vs. pH: ( a ) A I ; and (b) Fe 3÷. (Fig. 1 lc) which shows no obvioius relation-
ship with Br. Aluminum concentration does
line for seawater with a slope of <1 (0.85). increase with increasing degree of evaporation
This indicates that if seawater salts are the ma- (Fig. 1 ld), but the scatter in the data indicates
jor source for solutes in the system, then K is that pH may also be a controlling factor. As
removed early in the geochemical evolution of seen in Fig. 12a, AI concentration increases as
the water (Long et al., 1992 in this issue). This pH decreases. Fe (III) shows no clear relation-
appears to be a typical behavior for K in hy- ship with pH (Fig. 12b), suggesting that Fe
persaline environments including deep basin concentrations are controlled by a complex re-
brines (Gudramovics, 1981; Long and Gud- lationship among pH, salinity and oxidation-
ramovics, 1982; Wilson and Long, 1986; Wil- reduction processes.
son, 1989). The removal of K is thought to be
via sorption by clays early in the evolution of 4.4. Chemical modeling
the water (Jones et al., 1977). If sorption of K
occurs only in the early stages of brine evolu- Because of the limitations of the computer
tion, and saturation with K salts is not reached code and data set described earlier, with re-
during evaporation, then a plot of log~0 K vs. gard to aq)aeous activity coefficeints in con-
IOgl0 Br will have a slope of one (Carpenter, centrted brines, disequilibrium indices of alu-
1978). Because the slope is < 1 in this case, nite, jarosite and hydrous Fe-oxides could not
some process must continue to remove K from be directly calculated. Another approach to the
solution during evaporative concentration. study of mineral-water equilibrium is to con-
The removal of K from the concentrated so- struct activity-activity diagrams. For exam-
lutions could be explained by continued sorp- ple, Bladh (1982) studied theoretical solu-
tion of K during evaporation. However, the ef- tion-mineral equilibrium during the
ficiency of K sorption is decreased in weathering of sulfide minerals to produce ja-
hypersaline systems, particularly when diva- rosite, alunite and clays. Using the computer
lent cations are present, so continued removal code PATH (Helgeson et al., 1970 ) for the cal-
of K by this means during evaporation would culations, he plotted the changes in the loglo of
196 D.T. LONGET AL.

the activities of A l a + / ( H + )3 vs. F e 3 + / ( H + )3 one and have pH's < 4 (2.6-3.8). These data
during weathering. These changes could then are mainly from sipper and piezometer sam-
be compared to the alunite-jarosite phase ples taken along the TP line (Fig. 2). Samples
boundary which he defined with the reaction: which plot above this ratio are from the re-
gional groundwater with pH-value between 4.2
KAI3 ( S O 4 ) 2 ( O H ) 6 + 3 F e 3+ + 6H + and 6.2 and plots below the trend line of the
KFe3 (SO4)2(OH)6 + 3A13+ + 6 H + (1) first cluster. The trend of the data from the TP
line suggests possible solution equilibrium with
This phase boundary has a slope of one and alunite and jarosite.
water in equilibrium with both alunite and ja- In order to test this hypothesis, PHREEQE
rosite would plot along it. (Parkhurst et al., 1980) was used to predict to-
Because of the limitations discussed, activ- tal A13÷ and Fe 3÷ molalities and pH-values for
ity-activity diagrams could not be constructed solutions in equilibrium with alunite and ja-
for the study groundwater. Instead, a molal- rosite. The range of molalities necessary to de-
ity-molality plot of A13+/(H+) 3 vs. Fe3+/ fine a trend was calculated by varying the start-
(H ÷ )3 was constructed to detect a trend that ing pH of the modeled systems between 0.5 and
might indicate solution-mineral equilibrium. 4.0. The resulting model pH-values were al-
This diagram is shown in Fig. 13 and includes ways higher than the starting pH's. The results
all samples for which both A1 and Fe 3÷ were of the modeling are constrained by the choice
measured. The data appear to form a linear ar- of equilibrium constants (Keq) for the reac-
ray with a positive slope, but the slope is < 1. tions and by the chemistry of the solution being
This may be due to the fact that the data are modeled. Table 2 is a summary of selected
plotted as molalities rather than as activities or constants and appropriate reactions. The val-
that there is more than one data population in ues for amorphous Fe(OH)3 and goethite are
the diagram. from the WATEQ-PHREEQE data set (Ball et al.,
Closer inspection of the data shows that in- 1987; Parkhurst et al., 1987 ) and are generally
deed there are two trends on the diagram. The accepted as representative of these minerals
change in slope occurs at a log lo [ Fe 3+ / (H + ) 3 ] (D.K. Nordstrom, pers. commun., 1989).
ratio of 7, which corresponds to a pH of ~ 4.0 However, for alunite and jarosite there are a
(pH increases to the upper right on the dia- variety of equilibrium constants from which to
gram). Samples which plot below this ratio choose (Table 2).
cluster along a line with a slope that is nearly The differences in equilibrium constants for
alunite and jarosite are due to: (a) differences
15 . . . . . . . . N.~ ' " o' in the free energies assigned to these minerals
"
and in the thermodynamic data bases for
12 ~Xx
TP-Line • aqueous species used to calculate the equilib-
~ 9 rium constants, or (b) differences in solubility
+ t • Regional determinations or in their interpretation. For
6 • • • /\Groundwater
example, using the free energies of alunite and
o
,--1
3 jarosite from Kashkay et al. (1975) and ther-
modynamic data for aqueous species from
O I . . , . . , . , , . . , . .

either Naumov et al. (1971 ) or Robie et al.


0 3 6 9 12 15 ( 1978 ), log Keq ranges from - 12.69 to - 9.23
Log F e + 3 / ( H + ) 3 for jarosite and from - 3.62 to - 1.22 for alu-
Fig. 13. logjo[Al3+/(H+) 3] vs. lOglo[Fe3+/(H+) 3] with nit° (Table 2).
concentrations in molality. The equilibrium constants from Chapman et
ALUNITE, JAROSITE AND HYDROUS IRON OXIDES IN A HYPERSALINE SYSTEM 197

TABLE 2

Equilibrium constants (Iogto) considered in this study

Reference Jarosite* J Alunite .2 Amorphous Goethite *a


Fe(OH)3 .3

Hladky and Slanksy ( 1981 ) -9.08 - 1.66


Bladh ( 1982 ) - 7.12 - 1.54
Ball et al. (1979) - 14.80
Truesdell and Jones (1974) - 1_346 4.891
Nordstrom et al. (1989) - 1_0

Kashkay et al. (1975) *~ -9.23 - 1.87


Kashkay et al_ ( 1975 )'6 - 12.692 - 3.616
Chapman et al. (1983) - 9.21 - 1.22

*'KFej(SO4)2(OH)z+6H+~K+ + 3Fe3+ + 2SO24- +6H20.


*:AI3 (SO4)2 (OH)2 + 6 H + ~ K + + 3A13+ +2SO4:- +6H20.
*JFe (OH)3 + 3H + ~ F e 3+ + 3H20.
*4FeOOH + 3 H + ~ F e 3+ +2H20.
*SG°'s of alunite andjarosite from Kashkay et al. (1975), others from Naumov et al. ( 1971 ).
*6G°'s ofalunite and jarosite from Kashkay et al. (1975), others from Robie et al_ (1978).

al. ( 1983 ) were found to provide the best fit to these samples based on their similarity to sea-
the data. These constants are based on data water, degree of concentration and available
from N a u m o v et al. ( 1971 ) and Kashkay et al. data (Dickson et al., 1988; Long et al., 1992 in
( 1975 ) with modifications to the value for al- this issue).
unite from the work of Zotov (1971 ). These The result of the modeling is shown in Fig.
data represent an internally consistent ther- 14 as a solid line for halite disequilibrium of
modynamic data base. D.K. Nordstrom (pers.
commun., 1989), for example, has found this
data base and the constants for jarosite pre- 15 1 . . . . . . . . . . . . . . . .
sented by Chapman et al. ( 1983 ) to be better 12 /JAA . %[
in describing solution-mineral equilibria than • r :/ •
the other values listed in Table 2.
+
The modeling was done assuming a solution
in equilibrium with gypsum, although the
presence or absence of gypsum made little dif- ,-~ JAG JAROSITE
ference in the output. Disequilibrium
[logm(IAP/Ksp)] with respect to halite was 0 3 6 9 12 15
varied in the models between - 2 . 0 0 and 0.00,
Log Fe + 3 / ( H ~ 3
which is the range calculated for the samples
in Fig. 11 (Long et al., 1992 in this issue). Be- Fig. 14. Phase relationship of alunite and jarosite pre-
cause F is present in the groundwater (Dick- dicted from P H R E E Q E . Concentrations are in molality,
Solid line is theorectical molal phase boundary between
son et al., 1988) and F is known to form rela- alunite and jarosite in the presence of 6 ppm F - and a
tively stable aqueous complexes with both Fe disequilibrium index for halite of - 1. Dashed lines show
and A1 (Hem, 1968, 1985), solution-mineral expected range in position of phase boundary in spring
equilibrium was also studied as a function o f F zone because of variable F - concentrations and halite
saturation states. JAG and JAA points are the theoretical
concentration. Fluoride was added to the
equilibrium molal concentrations for jarosite-alunite-
model solutions as NaF ranging in concentra- goethite and jarosite-alunite-amorphous Fe(OH)~,
tion from 1 to 20 ppm, the expected range for respectively.
198 D.T. LONG ET AL.

- 1.00 and F concentrations of 6 ppm. This line either the JAG or the JAA point on the dia-
represents the molal phase boundary between gram suggests that neither hydrous Fe-oxide is
alunite and jarosite for an average sample from dominant in controlling the solution chemis-
the TP line. The dotted lines on either side of try. The spread of the data along the boundary
this line represent the range in the position of could be explained by variable mixtures of JAG
the boundary depending on halite disequili- and JAA solutions. It could also be explained
brium and F concentration. The samples from by solutions being in equilibrium with alunite,
the TP line cluster along this boundary, sup- jarosite and several hydrous Fe-oxides, each
porting the hypothesis that the groundwater in oxide having a different free energy. This might
the spring zone is in equilibrium with alunite be more realistic than mechanical mixtures of
and jarosite. The alunite-jarosite phase JAG and JAA solutions. The JAA and possibly
boundary is plotted as a dotted line at JAG would then represent end-member equi-
loglo[Fe3+/(H+) 3] ratios of >7, because libria in the system.
these minerals theoretically are not mutually There is scatter in the data about the phase
stable at pH's > 4 (Hladky and Slansky, 1981 ). boundary beyond the range attributable to the
The data which plot higher than the ratio of 7 variable F concentrations and disequilibrium
clearly do not cluster along the phase bound- states of halite considered in the modeling.
ary, which indicates that they are not con- Possible sources of data scatter include: (1)
trolled by alunite-jarosite equilibrium. The re- clean and analytical techniques; (2) use of 0.4-
suits suggest that alunite-jarosite equilibrium /~m filters rather than 0.1-/zm filters, which
controls the geochemistry of AI and Fe in the might affect the Fe and A1 data; ( 3 ) model so-
spring zone and that the spread of the data lution chemistries may not reflect natural con-
along the phase boundary is due to differences ditions; and (4) incomplete characterization
in pH among the samples. of the alunite and jarosite stoichiometries. We
However, the above models did not consider feel that our clean and analytical techniques
the presence of hydrous Fe-oxides in the sedi- were the best possible for this study (cf. Lyons
ments and the potential for the solution to be et al., 1992 in this issue). Filtration through
in equilibrium with alunite, jarosite and Fe- 0.1-#m filters was impossible under the cir-
oxide (Brown, 1971 ). Goethite and an amor- cumstances of this study, so this factor will
phous Fe-oxide were identified in the sedi- have to remain as a possible cause of the scat-
ments. Expected F e 3 + / ( H + ) 3 and A13+/ ter. The calculated range in the phase bound-
(H ÷ )3 ratios in solutions in equilibrium with ary in Fig. 14 does reflect the actual solution
alunite-jarosite-goethite (JAG) and alunite- conditions along the TP line. This is because
jarosite-amorphous Fe(OH)3 (JAA) equilib- the calculated phase boundaries considered the
ria were also calculated. The results of these m a x i m u m expected variation in the concen-
calculations are shown in Fig. 14. trations of C1 and F, the two most important
The equilibrium pH-values for the JAA so- ligands controlling Fe and AI concentrations,
lutions were higher ( ~ 3.80) than those for the respectively, at low pH.
JAG solutions ( ~ 1.30). Both showed little Incomplete characterization of the alunite
change when the starting pH was varied so that andjarosite could be a major cause of the scat-
the JAG and JAA models plot as tight clusters ter in the data. Although the XRD and EDS
in Fig. 14 rather than as lines. The JAA cluster data indicate that these minerals are close to
plots at the high end the data on the phase the K end-member, it is known that low-tem-
boundary. The JAG cluster plots on the phase perature alunites and jarosites can have vari-
boundary, but below the data. able amounts of hydronium (H3O+) substi-
The failure of the field data to cluster at tution for K, as dicussed by Alpers et al. ( 1992
ALUNITE, JAROSITE AND HYDROUS IRON OXIDES IN A HYPERSALINE SYSTEM 199

in this issue). This substitution, which is not Lake Tyrrell plot along a molal equilibrium
detected by EDS, would of course affect the free boundary between alunite and jarosite and be-
energies of these minerals, as would any solid tween equilibrium mixtures of alunite-jaros-
solution of Fe and A1 between the minerals. ite-goethite and alunite-jarosite-amorphous
Thus, resolution of the cause in the scatter Fe-oxide. The spread of the data along the
in the data requires better information on dis- boundary can be explained by solutions of
solved AI and Fe speciation as well as a better varying pH in equilibrium with alunite and ja-
understanding of the stoichiometry of alunite rosite and/or solutions in equilibrium with al-
and jarosite. Also, Pitzer's equations must be unite, jarosite and different hydrous Fe-ox-
employed in the chemical models for a more ides. In any case, the trend in the data when
rigorous test of solution-mineral equilibrium compared to the results of the chemical mod-
in the alunite-jarosite-Fe-oxide system. How- eling is consistent with the hypothesis that the
ever, PHREEQE (Parkhurst et al., 1980) did a groundwater in the spring zone is in equilib-
remarkable job in defining a phase boundary rium with alunite, jarosite, and possibly Fe-ox-
consistent with the field data. ides. This result is tentative because of the po-
tential limitations in the chemical modeling.
5. Summary However, considering these limitations, equi-
librium constants of 10 -~22 and 10 -9.23 ap-
Alunite and jarosite were investigated in the pear to best describe the hydrolysis reactions
hypersaline sediments of Lake Tyrrell, Aus- of alunite and jarosite, respectively. This sup-
tralia. These minerals occur as layers and ce- ports the results of Chapman et al, ( 1983 ) and
ments within the top 20 cm of the sediments. the work of Alpers et al. (1989) regarding the
The pore water in these sediments is highly formation ofjarosite in acid-mine drainages.
concentrated by evaporation and are precipi-
tating gypsum and halite. Alunite and jarosite Acknowledgements
appear in many cases to occur together as a
mechanical mixture within the sediments. This work was supported by NSF Grant
Their chemistry approximates the K end- EAR-12065 to D.T.L., W . B i . and M.E.H. We
member of the alunite and jarosite super- would like to thank Jean Long and Jane Matty
groups. Alunite occurs as equant grains of < 1 (Michigan State and Central Michigan Uni-
lzm in cross-section, while jarosite occurs as versities) for reviewing early versions of this
hexagonal platelets of > 1/zm in cross-section manuscript. Robert Lent (University of New
and ~ 2/zm thick. Hampshire) was instrumental in collecting and
We conclude that the occurrence of alunite the early processing of the field samples. Dis-
and j arosite in the sediments of the spring zone cussions with Charlie Alpers and Kirk Nords-
of Lake Tyrrell is consistent with a model in trom (U.S.G.S) were very important in our
which these minerals precipitate as a result of understanding of the geochemical nature of al-
evaporitic processes. The constituent ions are unite and jarosite. This work could not have
not necessarily locally derived. Groundwater been done without the help of our Australian
enters the playa as acid solutions with high colleagues. In particular we would like to thank
concentrations of K, SO4, A1 and Fe 3+. Within Andy Herczeg (CSIRO, Glen Osmond) and
the playa, these ions are concentrated by evap- Bruce Dickson (CSIRO, North Ryde).
oration. The geochemistry of A1 and Fe 3+is also
affected by pH and in the case of Fe 3+, oxida-
References
tion-reduction processes. Adams, F. and Hajek, B.F., 1978. Effectsof solution sul-
The field data from the spring zone area of fate, hydroxide,and potassium concentrationson the
200 D.T. LONGETAL.

crystallization of alunite, basaluminite, and gibbsite Garrels, R.M. and Thompson, M.E., 1962. A chemical
from dilute aluminum solutions. Soil Sci., 126: 169- model for seawater at 25 °C and one atmosphere total
173. pressure. Am. J. Sci., 260: 57-66.
Adams, F. and Rawajfih, Z., 1977. Basaluminite and alu- Gibbs, C.R., 1976. Characterization and application of
nite: a possible cause of sulfate retention by acid soils. Ferrozine iron reagent as a ferrous and total iron indi-
Soil Sci. Soc. Am. J., 41: 686-692. cator. Limnol. Oceanogr., 14: 357-367.
Alpers, C.N., Nordstrom, D.K. and Ball, J.W., 1989. Sol- Goldbery, R., 1978. Early diagenetic, nonhydrothermal
ubility ofjarosite solid solutions precipitated from acid Na-alunite in Jurassic flint clays, Makhtesh Ramon,
mine waters, Iron Mountain, California, U.S.A. Sci. Israel. Geol. Soc. Am. Bull., 89: 687-698.
Grol., 42: 281-298. Goldbery, R., 1980. Early diagenetic, Na-alunites in Mio-
Alpers, C.N., Rye, R.O., Nordstrom, D.K., White, L.D. cene algal mat intertidal facies, Ras Sudar, Sinai. Se-
and King, B.S., 1992. Chemical, crystallographic and dimentology, 27:189-198.
stable isotopic properties of alunite and jarosite from Gudramovics, R., 1981. A geochemical and hydrological
acid-hypersaline Australian lakes. In: W.B. Lyons, D.T. investigation of a modern coastal marine sabkha. Mas-
Long, A.L. Herczeg and M.E. Hines (Guest-Editors), ter's Thesis, Michigan State University, East Lansing,
The Geochemistry of Acid Groundwater Systems. Mich. (unpublished).
Chem. Geol., 96:203-226 (this special issue). Hartig, C., Brand, P. and Bohmhammel, K., 1984. Fe-A1
Bail, J.W., Nordstrom, D.K. and Zachmann, D.W., 1987_ lsomorphie und Strukturwasser in Kristallen von Ja-
WATEQ2 - - A personal computer FORTRAN trans- rosit-Alunit-Typ. Z. Anorg. Allg. Chem., 508:159-164.
lation of the geochemical model WATEQ2 with re- Harvie, C.E. and Weare, J.H., 1980. The prediction of
vised data base. U.S. Geol. Surv., Open-File Rep. 87- mineral solubilities in natural waters: the N a - K - M g -
50, 108 pp. Ca-C1-SO4-H20 system from zero to high concentra-
Barnes, R.B., 1975. The determination of specific forms tions at 25 degrees C. Geochim. Cosmochim. Acta, 44:
of aluminum in natural water. Chem. Geol., 15:177- 981-997.
191. Heigeson, H.C., Brown, T.H., Nigrini, A. and Jones, T.A.,
Bladh, K.W., 1982. The formation of goethite, jarosite, 1970. Calculation of mass transfer in geochemical pro-
and alunite during the weathering of sulfide-bearing cesses involving aqueous solutions. Geochim. Cos-
felsic rocks. Econ. Geol., 77: 176-184. mochim. Acta, 34: 569-592.
Bowler, J.M., 1986. Spatial variability and hydrologic Hem, J.D., 1968. Graphical methods of studies of aqueous
evolution of Australian lake basins: Analogue for aluminum hydroxide, fluoride, and sulfate complexes.
Pleistocene hydrologic change and evaporite forma- U.S. Geol. Surv., Water-Supply Pap. 1827-B, 33 pp.
tion. Palaeogeogr., Palaeoclimatol., Palaeoecol., 54:21- Hem, J.D., 1985. Study and interpretation of the chemi-
41. cal characteristics of natural water. U.S. Geol. Surv.,
Bowler, J.M. and Teller, J.T., 1986. Quaternary evaporite Water-Supply Pap. 2254, 263 pp. (3rd ed. )
and hydrological changes, Lake Tyrrell, north-west Hemley, J.J., Hostetler, P.B., Gude, A.J. and Mountjoy,
Victoria. Aust. J. Earth Sci., 33: 43-63. W.T., 1969. Some stability relations of alunite. Econ.
Brophy, G.P., Scott, E.S. and Snellgrove, R.A., 1962. Sul- Geol., 64: 599-612.
fate studies, II. Solid solution between alunite and ja- Hines, M.E., Lyons, W.B., Lent, R.M. and Long, D.T.,
rosite. Am. Mineral., 47:112-126. 1992. Sedimentary biogeochemistry of an acidic, sa-
Brown, J.B., 1971. Jarosite-goethite stabilities at 25 ° C, 1 line groundwater discharge zone in Lake Tyrrell, Vic-
atm. Miner. Deposita, 6: 245-252. toria, Australia. In: W.B. Lyons, D.T. Long, A.L. Her-
Carpenter, A.B., 1978. Origin and chemical evolution of czeg and M.E. Hines (Guest-Editors), The
brines in sedimentary basins. Okla. Geol. Surv., Circ., Geochemistry of Acid Groundwater Systems. Chem.
79: 60-77. Geol., 96:53-65 (this special issue).
Chapman, B.M., Jones, D.R. and Jung, R.F., 1983. Pro- Hladky, G. and Siansky, E., 1981. Stability of alunite
cesses controlling metal ion attenuation in acid mine minerals in aqueous solutions at normal temperature
drainage streams. Geochim. Cosmochim. Acta, 47: and pressure. Bull. Minrral., 104: 468-477.
1957-1973. Ivarson, K.C., Ross, G.J. and Mills, N.M., 1979. The mi-
Dickson, B.L., Giblin, A.M. and Herczeg, A.L., 1988. crobiological formation of basic ferric sulfates, II.
Geochemistry and radiochemistry of acid-saline waters Crystallization in presence of potassium, ammonium,
at Lake Tyrrell, Victoria. CSIRO (Commonw. Sci. Ind_ and sodium salts. J. Soil Sci. Soc. Am., 43: 908-912.
Res. Org. ), Inst. Miner. Energy Construct., Invest. Rep. Jones, B.F., Eugster, H.P. and Rettig, S.L., 1977. Hydro-
1771R, 47 pp_ geochemistry of the Lake Magadi Basin, Kenya. Geo-
Dutrizac, J.E. and Jambor, J.L., 1987. Behavior of cesium chim. Cosmochim. Acta, 41: 53-72.
and lithium during the precipitation of jarosite-type Kashkay, C.M_, Borovskaya, Yu.B. and Babazade, M.A.,
compounds. Hydrometallurgy, 17:251-265. 1975. Determination of Gf°29aof synthetic jarosite and
ALUNITE, JAROSITE AND HYDROUS IRON OXIDES IN A HYPERSALINE SYSTEM 201

its sulphate analogues, Geochem. Int., 12- 115-121. 1971. Handbook of Thermodynamic Data. Atomiz-
Knight, J.E., 1977. A thermochemical study of alunite, dat, Moscow, 373 pp. (translated by G.J. Soleimani,
enargite, luzonite, and tennantite deposits. Econ. Geol., U.S. Geol. Surv., NTIS/PB-226-722/AS, 1974; see es-
72: 1321-1336. pecially pp. 18 I - 186).
Lock, D.E., 1988, Alunite andjarosite formation in evap- Nordstrom, D.K., 1982. The effect of sulfate on alumi-
orative lakes of South Australia. SLEADS (Salt Lakes, num concentrations in natural waters: some stability
Evaporites, and Aeolian Deposits) 1988 Annu. Meet., relations in the system AI203-SO3-H20 at 298 K.
Abstr., pp. 48-51. Geochim. Cosmochim. Acta, 46:681-692.
Long, D.T. and Gudramovics, R_, 1982. Major element Parkhurst, D.L., Thorstenson, D.C. and Plummer, L.N.,
geochemistry of brines from the wind-tidal fiat area, 1980. PHREEQE - - A computer program for geo-
Laguna Madre, Texas, J. Sediment. Petrol., 53: 797- chemical calculations. U.S. Geol. Sum., Water-Re-
810. sour_ Invest. 80-96, 194 pp. (revised and reprinted Jan.
Long, D.T., Fegan, N,E., Lyons, W.B., Hines, M.E., Mac- 1985).
umber, P.G. and Giblin, A.M., 1992. Geochemistry of Pitzer, K.S., 1979. Theory: ion interaction approach. In:
acid brines: Lake Tyrrell, Victoria, Australia. In: W.B. R.M. Pytkowicz (Editor), Activity Coefficients in
Lyons, D.T. Long, A.L. Herczeg and M.E. Hines Electrolyte Solutions, Vol. 1. CRC Press, Boca Raton,
(Guest-Editors), The Geochemistry of Acid Ground- Fla., pp. 157-208
water Systems. Chem. Geol., 96:33-52 (this special Plummer, L.N., Parkhurst, D.L., Fleming, G.W. and
issue)_ Dunkle, S.A., 1988. P H R Q P I T Z - A computer pro-
Lyons, W.B., Chivas, A.R., Lent, R.M., Welsh, S., Kiss, gram for geochemical calculations in brines. U.S. Geol.
E, Mayewski, P.A., Long, D.T. and Carey, A.E., 1989. Surv., Water-Resour. Invest. Rep. 88-4153, 310 pp.
Metal concentrations in surficial sediments from hy- Presley, F_B., 1971. Determination of selected minor and
persaline lakes, Australia. Hydrobiologia, 197:13-22. major inorganic constituents_ In: E.L. Winterer, W.R.
Lyons, W.B., Welch, S., Long, D.T., Hines, M.E., Giblin, Riedel, et al. (Editors), Initial Reports of the Deep Sea
A.M., Carey, A.E, Macumber, P.G., Lent, R.M. and Drilling Project, Vol. VII. U.S. Gov. Print_ Off-, Wash-
Herczeg, A.L., 1992. The trace-metal geochemistry of ington, D.C., pp. 1749-1755.
the Lake Tyrrell system brines (Victoria, Australia)_ Reardon, E.J. and Beckie, R.D., 1987. Modeling chemical
In: W.B. Lyons, D_T. Long, A.L. Herczeg and M.E. equilibria of acid mine-drainage: the F e S O 4 - H 2 S O 4-
Hines (Guest-Editors), The Geochemistry of Acid H20 system. Geochim. Cosmochim. Acta, 51: 2355-
Groundwater Systems. Chem. Geol., 96:115-132 (this 2368.
special issue). Robie, R.A., Hemingway, B.S. and Fisher, J.R., 1978.
MacCaffrey, M.A., Lazar, B. and Holland, H.D., 1988. The Thermodynamic properties of minerals and related
evaporation path of seawater and the coprecipitation substances at 298_ 15 K and 1 bar (105) pascals) pres-
of Br and K with halite, (Unpublished.) sure and at higher temperatures. U.S. Geol. Surv., Bull.
Mackin, J.E. and Aller, R.C_, 1984. Dissolved AI in sedi- 1452, 456 pp.
ments and waters of the East China Sea: Implications Rye, R.O., Bethke, P.M. and Wasserman, M.D., 1989.
for authigenic mineral formation. Geochlm. Cosmo- Diverse origins of alunite and acid-sulfate alteration:
chim. Acta, 40: 218-297. stable isotope systematics. U.S. Geol. Surv., Open-File
Macumber, P.B., 1983. Interactions between ground water Rep. 89-5, 18 pp.
and surface systems in northern Victoria. Ph.D. Dis- Sass, E., Nathan, Y. and Nissenbaum, A., 1965. Mineral-
sertation, University of Melbourne, Melbourne, Vic., ogy of certain pyrite concretions from Israel and their
506 pp. alteration products. Mineral. Mag., 35: 84-87.
Macumber, P.G., 1992. Hydrological processes in the Scott, K.M., 1987. Solid solution in, and classification of,
Tyrrell Basin, southeastern Australia. In: W.B. Lyons, gossan-derived members of the alunite-jarosite fam-
D.T. Long, A.L Herczeg and M.E. Hines (Guest-Edi- ily, northwest Queensland, Australia. Am. Mineral., 72:
tors), The Geochemistry of Acid Groundwater Sys- 178-187.
tems. Chem. Geol., 96:1-18 (this special issue). Stoffregen, R_E. and Alpers, C.N., 1987. Woodhouseite
Mann, A.W., 1983. Hydrogeochemistry and weathering and svanbergite in hydrothermai ore deposits: prod-
on the Yilgarn Block, W.A. - - Ferrolysis and heavy ucts of apatite destruction during advanced arglllic al-
metals in continental brines. Geochim. Cosmochim. teration. Can. Mineral., 25:201-211,
Acta, 47: 181-190. Stookey, L.L., 1970. Ferrozine - - a new spectrophoto-
Michel, F.A_ and van Everdingen, R.O., 1987. Formation metric reagent for iron. Anal. Chem., 42:779-781.
of a jarosite deposit on Cretaceous shales in the Fort Sullivan, P.J., Mattigod, S.V. and Sobek, A.A., 1986. Dis-
Norman area, Northwest Territories. Can. Mineral., 25: solution of iron sulfates from pyritic coal waste. Envi-
221-226_ ron. Sci. Technol., 20: 1013-1016.
Naumov, G.B., Ryzhenko, B.N. and Khodakovsky, I.L., Teller, J.T., Bowler, J_M. and Macumber, P_G., 1982.
202 D.T. LONG ET AL.

Modem sedimentation and hydrology in Lake Tyrrell, the Michigan Basin brine. Ph.D. Dissertation, Michi-
Victoria. J. Geol. Soc. Aust., 29: 159-175. gan State University, East Lansing, Mich.
Truesdell, A.H. and Jones, B.F., 1974. W A T E Q - A com- (unpublished).
puter program for calculating chemical equilibria of Wilson, T.P. and Long, D_T., 1986. Constraints on the
natural waters. J. Res. U.S. Geol. Surv., 2: 233-248. evolution of the Michigan Basin brines. Geol. Soc. Am.,
van Breemen, N., 1973. Dissolved aluminum in acid sul- Abstr. Prog., 18: 791.
fate soils and in acid mine waters_ Soil Sci. Soc. Am. Zotov, A.V., 1971. Dependence of the composition of al-
Proc., 37: 694-697. unite on the temperature of its formation. Geochem.
Wilson, T.P., 1989_ Origin and geochemical evolution of Int., 8: 71-75_

View publication stats

You might also like