You are on page 1of 25

0361-0128/10/3885/509-25 509

Sedimentary and Geobiological Perspectives


on Sulfur Cycle Evolution
SEDIMENTOLOGISTS, geochemists, and geobiologists have long
tried to address critical questions about sulfur cycle evolution.
What is the origin of oceanic sulfate? What is the timing of the
onset of bacterial sulfate reduction? How are sulfur species
transformed and transferred between different reservoirs?
Some workers have used geologic approaches to trace evidence
of formation of sulfur-bearing minerals in sedimentary settings
(e.g., Grotzinger and Kasting, 1993; Pope and Grotzinger,
2003; Schrder et al., 2008). Others have developed methods
that link sulfur, iron, oxygen, and trace metal geochemistry
(e.g., Anbar and Knoll, 2002; Algeo and Lyons, 2006; Canfield
et al., 2008; Scott et al., 2008). Still others have calibrated sea-
water geochemistry and sulfate concentrations using fluid in-
clusions (e.g., Horita et al., 1991, 1996, 2002; Lowenstein et
al., 2005). Some have used biological methods to document
the presence of different types of sulfur-utilizing organisms in
ancient environments (Brocks and Schaeffer, 2008), and oth-
ers have developed and applied methods that rely on mea-
surements of the stable isotopes of sulfur and oxygen to trace
changes in the cycling of sulfur and the types of transforma-
tions that occurred at different times in Earth history (Mon-
ster et al., 1979; Cameron, 1982, 1983; Canfield, 2001).
Changes in the sulfur cycle have been attributed to the evo-
lution of Earth surface chemistry, evolving chemistry of the
atmosphere and oceans, and to accompanying changes in
ecology of sulfur-utilizing organisms. There is a generally, but
not universally, accepted model for sulfur cycle evolution.
This model includes an early stage (Archean and earliest Pro-
terozoic) with low sulfate oceans (less than 100200 M), a
component of inferred atmospheric origin that reflects low
oxygen content, and a small role for sulfate reduction. This
model calls for a change in the sulfur cycle to slightly higher
oceanic sulfate concentrations (millimolar) as a result of greater
production of sulfate by oxidative weathering of continental
sulfide starting at ~2.4 Ga and extending for almost the entire
Proterozoic. This change also included more significant recy-
cling of sulfide by oxidative weathering and ultimately led to
a change in ocean chemistry by the Mesoproterozoic. A con-
sensus on the sulfur cycle in the Mesoproterozoic has not
been reached because there appears to be evidence for sul-
fidic conditions in some settings and evidence for nonsulfidic
conditions in others. Debate presently focuses on the extent
of sulfidic and nonsulfidic ocean domains. Sulfate concentra-
tions are thought to have risen to levels of 20 to 30 mMby the
end of the Proterozoic (present-day concentrations), but to
have declined twice in the Phanerozoic to levels of a few mM
(in the latest Cambrian to Early Devonian, and in the Juras-
sic). Nevertheless, sulfate levels throughout the Phanerozoic
were higher than those in the Proterozoic. The controls on
sulfate levels reflect the supply of sulfate via oxidative weath-
ering and the efficiency of pyrite burial, and possibly organic
sulfur, as a sulfide sink.
The Isotopic Record of Sulfur Cycle Evolution
The isotopic record is the most continuous archive of the evo-
lution of the sedimentary sulfur cycle. Historically, it has fo-
cused on the study of
34
S (
34
S = (
34
S/
32
S)
sample
/(
34
S/
32
S)
reference
1). All of these values are given in units of per mil () and
the factor 1,000 is, therefore, not included in the definition.
More recently it has been recognized that additional infor-
mation can be provided by also considering variations among
other sulfur isotopes, particularly using
33
S, and
36
S (
33
S
Connections between Sulfur Cycle Evolution, Sulfur Isotopes, Sediments,
and Base Metal Sulfide Deposits
JAMES FARQUHAR,
1,2,
NANPING WU,
1
DONALD E. CANFIELD,
2
AND HARRY ODURO
1
1
Department of Geology and Earth System Science Interdisciplinary Center (ESSIC), University of Maryland,
College Park, Maryland 20740
2
NordCEE and the Institute of Biology, Syddansk Universitetet, 5020 Odense C., Denmark
Abstract
Significant links exist between the sulfur cycle, sulfur geochemistry of sedimentary systems, and ore deposits
over the course of Earth history. A picture emerges of an Archean and Paleoproterozoic stage of the sulfur cycle
that has much lower levels of sulfate (<200 M), carries a signal of mass-independent sulfur, and preserves
evidence for temporal and spatial heterogeneity that reflects lower amounts of sulfur cycling than today. A sec-
ond stage of ocean chemistry in the Paleoproterozoic, with higher atmospheric oxygen and oceanic sulfate at
low millimolar levels, follows this stage. The isotopic record in sedimentary rocks and in sulfide-bearing ore
deposits suggests abundant pyrite burial and implies a missing
34
S-depleted pool that may have been lost via
deep ocean deposition and possibly subduction. Proterozoic ocean chemistry appears to be quite complex. The
surface waters of the Proterozoic oceans are believed to have been oxygenated, but geologic evidence from ore
deposits and sedimentary rocks supports coexistence of significant sulfidic and nonsulfidic, anoxic, interme -
diate water and deep-water pools in the Mesoproterozoic. This stage in ocean chemistry ends with the second
major global oxidation event in the latest Neoproterozoic (~600 Ma). This event started the transition to more
oxygenated intermediate and deep waters, and higher but variable oceanic sulfate concentrations. The event
set the scene for the formation in the Phanerozoic of the first significant MVT deposits and possibly is reflected
in changes in other sedimentary rock-hosted base metal sulfide deposits.

Corresponding author: e-mail, jfarquha@Glue.umd.edu
2010 Society of Economic Geologists, Inc.
Economic Geology, v. 105, pp. 509533
Submitted: March 16, 2009
Accepted: February 5, 2010
= (
33
S/
32
S)
sample
/(
33
S/
32
S)
reference
-[(
34
S/
32
S)
sample
/(
34
S/
32
S)
reference
]
0.515
and
36
S = (
36
S/
32
S)
sample
/(
36
S/
32
S)
reference
-[(
34
S/
32
S)
sample
/(
34
S/
32
S)
reference
]
1.9
; see Appendix 1 for information on mass-depen-
dent and mass-independent sulfur isotope geochemistry).
Isotopic record before ~2400 Ma
The historical interpretations of
34
S of sedimentary pyrite
and sedimentary sulfate were made using a record similar to
the updated version presented in Figure 1. It was recognized
that this record preserves variations in the range of
34
S, the
difference between the seawater sulfate
34
S and the minima
of pyrite
34
S, over geologic time (cf. Monster et al., 1979;
Cameron, 1982, 1983; Canfield, 2001). One of the times
where a change was noted is approximately 2400 Ma. The
mean
34
S of sedimentary pyrite before 2400 Ma is approxi-
mately 2 per mil and the range of variability is small (Fig. 1B),
suggesting high degrees of pyrite burial and a diminished role
for processes that produce highly fractionated sulfide sulfur.
The change at ~2400 Ma (Fig. 1) is attributed to an ecosys-
tem response that is related to how metabolisms of sulfate-re-
ducing bacteria respond to different sulfate concentrations
(e.g., Cameron, 1982, 1983; Canfield, 2001). Early work
(Harrison and Thode, 1958) demonstrated that the discrimi-
nation among sulfur isotopes generated during the process of
sulfate reduction by bacteria was smaller when sulfate con-
centrations were very low. More recent work has confirmed
this and has added a dimension to the interpretation of the
change at 2400 Ma by identifying and calibrating a depen-
dence on ecosystem level expression of these isotope fraction-
ations that results from changes in sulfur transport and sulfate
reduction rates within unbioturbated sediment at low sulfate
concentrations (Canfield et al., 2000; Habicht et al., 2002).
The results of Habicht et al. (2002) suggest that sulfate trans-
port within the sediment is correlated with sulfate concentra-
tions in the overlying water column. This limits the preserva-
tion of the fractionations produced by sulfate-reducing
bacteria in sedimentary pyrite and in the geologic record, plac-
ing limits on oceanic sulfate concentrations at less than 200
M before ~2.4 Ga, and greater than 200 M after this time.
The record of
33
S and
36
S provide additional insight into
the evolution of the sulfur cycle at this time. The variations of

33
S and
36
S in rocks older than ~2.42 Ga have been attrib-
uted to mass-independent frationation of sulfur isotopes
(MIF-S) in atmospheric source reactions. This signal has
been interpreted to reflect atmospheric and surface weather-
ing environments with significantly lower oxygen levels for
the principal reason that it is a prerequisite for the produc-
tion, transfer, and preservation of MIF-S in surface sulfur
pools (Farquhar et al., 2000, 2007a; Kasting, 2001; Farquhar
and Wing, 2003, 2005). These observations support a model
of a pre-2.42 Ga sulfur cycle with a diminished role for bio-
logical reduction of sulfate and for diminished oxidation of
sulfide because active cycling would be a process that would
homogenize and remove the MIF-S signal (Farquhar et al.,
510 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 510
-60
-40
-20
0
20
40
60
80
0 500 1000 1500 2000 2500 3000 3500 4000
Age (Ma)

3
4
S
Literature data
Pyrite (avg)
Seawater sulfate (est)
A
50
150
250
350
450
0
200
400
50
150
250
350
50
150
250
50
150
250
sulfide means
X = -16.9+/- 1.2
0-540 Ma
X = 7.2 +/- 1.6
540-900 Ma
X = 3.9 +/- 1.2
900-1850 Ma
X = 1.7 +/- 0.4
1850-2400 Ma
X = 1.6 +/- 0.4
2400-3900 Ma
B.
n
u
m
b
e
r
n
u
m
b
e
r
n
u
m
b
e
r
n
u
m
b
e
r
n
u
m
b
e
r
-60 -30 0 30 60

34
S
FIG. 1. A. Compilation of
34
S vs. time gray circle symbols (Canfield, 1998; Prokoph et al., 2007; Canfield and Farquhar,
2009) with running average for sulfide (gray line) and inference about seawater sulfate (black line) that we fit using a cubic
spline to sulfate data younger than 2450 Ma and constrained by literature estimates for periods prior to 2450 Ma. The curve
is similar to previous curves for the Phanerozoic. The positive values for seawater sulfate and average sulfide indicate an im-
balance in the sulfur cycle record extending for almost 2000 m.y. until ~500 Ma. B. Histograms of the frequency of obser-
vations of sulfide (gray filled bars), and sulfate from the black curve (hatched bars), for five intervals in geologic time. The
height of each histogram is normalized to the total number of samples in order to put them on a similar aspect ratio. Means
for sulfide data are also plotted. Uncertainties on the mean are determined using a monte carlo resampling technique.
2000). Table 1 synthesizes observations from the integrated
record of sulfur isotopes (
34
S,
33
S, and
36
S), as well as pro-
viding implications from these observations. There are sev-
eral issues with the characteristics of the MIF-S signal that
are important in shaping our understanding of the cycling of
sulfur in the Archean because they relate to the controls on
the manifestation of the
33
S (
34
S and
36
S) signal in the
sedimentary rock record.
First, there is a change in the magnitude of the range of

33
S (
34
S and
36
S) over the course of the Archean. There is
more significant variability for
33
S in the Neoarchean, a
damped
33
S signal in the Mesoarchean, and an intermediate
magnitude
33
S signal in the Paleoarchean (Fig. 2).
Second, there is clear evidence for small-scale high fre-
quency variations in sedimentary rocks, mostly shales, at the
thin-section and hand-sample scale, but there are also indica-
tions that underlying longer timescale correlations exist be-
tween cores of roughly equivalent successions. Profiles are
collected for three roughly equivalent sections through the
Naute Shale of the Transvaal Supergroup and two successions
through the equivalent Australian Mount McRae Shale (Fig.
3). These successions can be roughly correlated using two
marker beds: Brunos band in the deeper parts and the tran-
sition to Kuruman and Brockman banded iron formation in
the upper parts. These data suggest the presence of positive
excursions for
33
S in samples from the lower parts of these
successions. Variations in the correlation between
34
S and

33
S (and
36
S) also appear to be shared between these sec-
tions (Kaufman et al., 2007; Ono et al., 2009b; Fig. 4).
Third, a number of studies that used microanalysis of sulfur
isotopes (Mojzsis et al., 2003; Papineau et al., 2005, 2007; Cates
and Mojzsis, 2006; Kamber and Whitehouse, 2007; Philippot
et al., 2007; Ono et al., 2009a) have demonstrated homo-
geneity rather than heterogeneity for
33
S on a thin-section
scale. Of the studies that demonstrated heterogeneity, there
appears to be a relationship between variations for
33
S and
the type of pyrite that is present in the thin sections. For two
studies of the same late Neoarchean successions described
above from the Naute Shale, evidence of significant hetero-
geneity was also demonstrated (Kamber and Whitehouse,
2007; Ono et al., 2009a). The fine-grained pyrite preserved
the most significant
33
S signals, whereas larger spheroidal
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 511
0361-0128/98/000/000-00 $6.00 511
TABLE 1. Observations and Implications from the Multiple Sulfur Isotope Record
Older than ~2.42 Ga
The presence of Archean MIF-S implies processes (photochemical) in addition to microbial sulfate reduction played an important role in the sulfur
cycle
Covariation between
34
S,
33
S, and
36
S tells us that sulfur cycle processes were not efficient enough to homogenize MIF-S once it was present
Grain scale to thin section scale heterogeneity for
33
S (e.g., Ono et al., 2003; Kamber and Whitehouse, 2007) implies mixing of sulfur with different
MIF-S compositions, suggesting more than one sulfur pool (e.g., sulfate and sulfide-polysulfides-sulfur) or short residence times (low concentrations) for
sulfur
Coherent relationships between
36
S and
33
S at the formation level (Fig. 4) (e.g., Farquhar et al., 2007; Kaufman et al., 2007) imply gross changes in
MIF-S source reactions and/or ecology that vary in time and possibly with geography
Observation of MIF-S signals in hydrothermal VMS systems (Farquhar and Wing, 2005; Jamieson et al., 2006) and in marine sediments subject to diage-
netic processes implies the existence of multiple MIF-S sulfur cycle pools
A fractionation of ~20 is observed for
34
S of the most
34
S-depleted sulfides and seawater sulfate (Fig. 1)
~2.42 Ga to ~600 Ma
Samples of the ~2.42 Ga Duitschland Fm. capture the transition from a MIF-S world to a MDF-S world (Bekker et al., 2004; Guo et al., 2009)
The transition from a MIF-S world to a MDF-S world appears to be correlated with one of the Early Proterozoic carbon isotope excursions
The average
34
S of Proterozoic sedimentary pyrite appears to be near zero, (if not slightly positive) and implies high fractions of pyrite burial and a
missing
34
S-depleted sulfur pool
Studies of basins have shown more negative
34
S is observed in deeper parts of sedimentary basins (Shen et al., 2002) implying the missing
34
S-depleted
sulfur pool may have formed and been buried in deeper water settings (Logan et al., 1995), possibly subducted (Canfield, 2004)
Deep- to shallow-water variation in the
34
S of the pyrite pool may reflect distillation of sulfur by sulfate reduction in Proterozoic oceans, reflecting low
sulfate concentrations (Johnston et al., 2006)
A relationship between
34
S and
33
S is seen in some sections that implies an isotopic reservoir (Rayleigh) effect (Johnston et al., 2005, 2006)
Significant (short time scale, high frequency) variations of
34
S for carbonate associated sulfate (a proxy for seawater sulfate) have been interpreted to in-
dicate low (mM) sulfate concentrations (Kah et al., 2004; Lyons et al., 2004)
The
33
S of seawater sulfate rose during the middle Proterozoic (between ~1.6 and ~1.3 Ga) (Johnston et al., 2005a) and this observation is interpreted
to be related to a change in the ecology of the ocean-sediment system to favor expression of fractionations of sulfur disproportionators
A fractionation of ~40 is observed for
34
S of the most
34
S-depleted sulfides and seawater sulfate (Fig. 1)
Since ~600 Ma
The variability with time of
34
S for seawater sulfate generally diminishes (e.g., Kampschulte and Strauss, 2004; Claypool et al. 1980), but some evidence
for higher frequency oscillations are still present (e.g., Paytan et al., 1998, 2004; Kurtz et al., 2003; Lyons et al., 2004); this has been interpreted to re-
flect growth of the oceanic sulfate pool to a size large enough to dampen, but not completely eliminate, the variability in
34
S caused by changes in the
rate of or fractionations associated with pyrite burial
A fractionation of ~60 is observed for
34
S between the most
34
S-depleted sulfides and seawater sulfate (Fig. 1)
The magnitude of the fractionation between buried pyrite and sulfate appear to have changed from ~25 before ~550 Ma to ~35 between ~550 and
~300 Ma before increasing to ~43 by about 250 Ma (Wu et al., 2010)
512 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 512
-4
-2
0
2
4
6
8
10
12
0 1000 2000 3000 4000
Sample Age (Millions of years)

3
3
S
Younger than 2450 Ma
Older than 2450 Ma
Neo Meso Paleo Eo
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0 500 1000 1500 2000
Age (Mya)

3
3
S
FIG. 2. Plot of
33
S versus sample age from a compilation of values reported in (Farquhar et al., 2000, 2002, 2007a; Hu
et al., 2003; Mojzsis et al., 2003; Ono et al., 2003, 2006a, 2007; Bekker et al., 2004; Farquhar and Wing, 2005; Johnston et
al., 2005a, 2006, 2008; Papineau et al., 2005, 2007; Cates and Mojzsis, 2006; Ohmoto et al., 2006; Kamber and Whitehouse,
2007; Domagol-Goldman et al., 2008; Partridge et al., 2008; Thomazo et al., 2009). Gray filled circle symbols are samples
older than 2.45 Ga and white filled circle symbols are for samples younger than 2.45 Ga. Ancient samples reflect mass inde-
pendent reactions and preservation of isotopic signals in the rock record. Samples younger than 2.45 Ga are interpreted to
carry a mass conservation-type signal associated with mass-dependent fractionation by BSR and other sulfur cycle processes
(see inset for sulfate).
270
280
290
300
310
320
330
A B C D E
245
255
265
275
285
295
305
315
325
335
D
e
p
t
h

(
m
)
100
110
120
130
140
150
160
170
180
190
225
235
245
255
265
275
285
295
305
315
165
175
185
195
205
215
225
235
245
-5 0 5 10 15 -5 0 5 10 15

33
S
33
S
-5 0 5 10 15

33
S
-5 0 5 10 15

33
S
33
S
-5 0 5 10 15
FIG. 3. Plot of
33
S as a function of depth in five drill cores that intersect roughly equivalent successions of the Naute
Shale (A, B, C GKP01, GKF01, and AD-5) and the Mount McRae Formation (D, E ABDP-9 and AUS 493). The upper
dark line represents the top of the formation and is taken as a time boundary. The lower dark line represents the bottom of
the formation (corresponding to its contact with a BIF horizon referred to as Brunos band). Where the lines (bands) are gray
instead of black, the position of this horizon is more uncertain. For panel C (AD-5) the top of the gray band represents the
bottom of the drill hole, and the bottom of the gray band is 78 m lower than the top of the formation which is similar to that
seen in cores GKP01 and GKF01. The data and the positions of the horizons are replotted from Ono et al. (2003, 2009b)
using information about the thickness of the Mount McRae Formation and Brunos band from Kakegawa et al. (1998) and
information about the GKP01 and GKF01 cores from Knoll and Beukes (2009) and Kaufman et al. (2007) using information
from Beukes et al. (1990) for the stratigraphy of AD-5 core.
grains and concentrations of pyrite in the same sections pre-
served near zero or even negative
33
S. These variations were
interpreted to reflect different generations of pyrite and for-
mation in a gradient, or stratified, system, including a strati-
fied water column (Kamber and Whitehouse, 2007).
A number of models have been proposed to explain
33
S
variations in the geologic record. They include suggestions
that the magnitude reflects variations in the proportion of at-
mospheric sulfur with a much larger initial
33
S (e.g., Far-
quhar et al., 2000, 2001), variations in
33
S of the atmos-
pheric components and therefore directly reflect a change in
the atmospheric source reactions (Ono et al., 2003; Domagol-
Goldman et al., 2008; Farquhar et al., 2007a), or the nature
and residence time of the sulfur pools in early Earth environ-
ments (Farquhar and Wing, 2005; Ono et al., 2009a, b). It
would appear that components of each of these suggestions
would be required to account for the three observations
above. A hybrid model that describes controls on the sulfur
cycle is thus favored (Fig. 5).
A recurring question in the literature is whether there is an
underlying systematic variation in the magnitude of the mass-
independent signal with respect to locality and or time, an
important issue for understanding the connections with the
systematics of sulfur in Archean ore deposits. There is clear
evidence for heterogeneity on a hand-sample scale (e.g.,
Kamber and Whitehouse, 2007; Ono et al., 2009a), but there
are also similarities in the magnitude of the mass-indepen-
dent signals that appear to be preserved at least on a basin
scale (Fig. 3). Resolution of this issue may be provided in part
by considering the model (Fig. 5) in the context of pyrite for-
mation and the atmospheric lifetime of volcanic sulfur.
Pyrite formation and burial in the present-day sulfur cycle
is not uniformly distributed in oceanic sediments. Estimates
of sulfate reduction rates suggest that approximately 70 to 80
percent of the sulfate reduction occurs at depths less than 200
m (Turchyn and Schrag, 2004, 2006; Canfield et al., 2005).
This number climbs to greater than 90 percent for depths less
than 800 m, which represent approximately 40 percent of the
total ocean area. Pyrite formation fixes sulfur ultimately de-
rived from the sulfate pool in these areas. In the Archean,
pyrite may also have acquired part of its signal from elemen-
tal sulfur that is inferred to have had large
33
S enrichments,
but there may be a problem with invoking elemental sulfur as
the source of the large positive
33
S values for pyrite.
Whereas sulfate may accumulate in the oceans and be deliv-
ered from seawater to the loci of pyrite formation, no accu-
mulation mechanism has been described for elemental sulfur.
Such a mechanism is needed to establish a distinct and large
standing pool of elemental sulfur for pyrite formation. It is
also possible that the deposition of elemental sulfur was not
uniform on a global scale, and was instead controlled by the
locations of volcanic and perhaps biogenic sources of atmos-
pheric sulfur gases, atmospheric residence times, and atmos-
pheric transport. The second possibility is supported by
calculations of the lifetime of sulfur dioxide before it is pho-
tolyzed, which for sulfur gases in a low oxygen atmosphere is
less than one day (Farquhar et al., 2001). This is short com-
pared to the global atmospheric timescales for mixing and
rainout. If this were the case, then local deposition rates for
atmospheric sulfur compounds without significant positive

33
S values might come within an order of magnitude of the
reduced sulfur produced by sulfate reduction in regions prox-
imal to volcanic sources. The extent of these regions is not
known, but evidence to suggest that they may have been quite
large comes from the general structure of
33
S records for the
Hamersley, Transvaal, and Griqualand West basins (Fig. 3).
The change in the range of
33
S through the Archean also,
in part, reflects a change in the local fluxes of sulfur gases.
The reason for such a change is unclear. It needs to be deter-
mined whether large changes can be seen outside of the
Hamersley, Transvaal, and Griqualand West basins. It may re-
flect changes in volcanic fluxes, or a new contribution from
biogenic reduced sulfur gases, or possibly a change in the sul-
fur content of the volcanic gases resulting from the develop-
ment of a pool that is recycled through subduction. The dif-
ference between these Neoarchean signals and signals that
preceded them remains an important focus for future work.
An alternative possibility (Watanabe et al., 2009) suggests
that reactions other than those that occur in a gas phase may
produce anomalous sulfur isotope signals. These authors pre-
sent the results of sulfate reduction experiments using various
amino acids at temperatures between 150 and 200C. They
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 513
0361-0128/98/000/000-00 $6.00 513
-15
-10
-5
0
5
-5 0 5 10 15

33
S

3
6
S
Mt McRae Upper
Mt McRae Lower
Gamohaan Upper
Gamohaan Lower
FIG. 4. Plot of
33
S versus
36
S for pyrite samples from the Neoarchean
of South Africa (Gamohaan Fm.) and Western Australia (Mt McRae Fm.)
(Kaufman et al., 2007). These two successions are considered to be time
equivalent and can be correlated via marker beds. The upper and lower parts
of the two formations exhibit similar changes in the relationship between

33
S and
36
S. The lower parts of these successions are more linear, follow-
ing
36
S ~ 0.85
33
S, and the upper portions of these successions show more
scatter and
36
S >0.85
33
S. Slopes on lines in plot are 0.85 and 1.37. This
difference has been interpreted to reflect a change in mass independent
source reactions and also an overprint by cycling of sulfur in the water col-
umn in the upper parts of these two successions (Kaufman et al., 2007). It has
recently been suggested that the upper interval of the Mount McRae For-
mation may reflect local sulfidic conditions (Reinhard et al., 2009). In this
context, it is possible that a significant part of the change in
33
S vs.
36
S may
reflect stronger expression of sulfur cycling by sulfate reducers in a sulfidic
water column. This would not require a change in sulfate concentration, but
would imply more efficient expression of fractionations associated with sul-
fate reduction in the water column and would also be consistent with asser-
tions for water column sulfate reduction by Kamber and Whitehouse (2007).
suggest that mass-independent fractionation under such con-
ditions may involve a magnetic isotope effect (MIE) (e.g.,
Turro, 1983; Buchachenko, 1995: see App. 1) or a surface cat-
alyzed reaction (e.g., Lasaga et al., 2008), but further work
will be needed to resolve this. This proposal might be taken
further to suggest that atmospheric reactions are not the
source of the mass-independent signal, and that mass-inde-
pendent fractionation (MIF) is a result of processes that
occur in sediments. We do not think that this is presently a
defendable suggestion because several issues need to be ad-
dressed relating to the following: (1) the nature of the signal,
such as the tight coupling of
33
S and
36
S in many Archean
samples (Fig. 4); (2) the production and transfer of the signal
through the wide range of Archean sedimentary, metamor-
phic, and igneous rocks; and (3) the apparent geographic and
temporal coherence of the signal over the course of the
Archean and its abrupt disappearance in the Paleoprotero-
zoic. Anomalous sulfur isotope fractionation arising from such
reactions may be relevant in some natural environments
where thermochemical sulfate reduction occurs, although ev-
idence has yet to be presented that this is the case.
Isotopic record since ~2400 Ma
The sulfur cycle since ~2400 Ma is characterized by a larger
range in the signal of
34
S and a lack of significant mass-inde-
pendent signals in sedimentary rocks, suggesting an end to
the period of MIF-S and a transition to a sulfur cycle more
strongly influenced by biological processes (Bekker et al.,
2004; Papineau et al., 2005; Guo et al., 2009). The upper
parts of the Duitschland Formation, South Africa, de-
posited between 2450 and 2320 Ma, capture this transition
for both
34
S and
33
S. They also show a correspondence
between
34
S,
33
S, and
13
C, which reinforces inferences
about a link between the carbon and sulfur cycles at the
time of the rise of oxygen. High
13
C values for carbonates
reflect carbon burial events that were associated with sig-
nificant oxygen production (Karhu and Holland, 1996). The
coincidence of these signals links sulfate concentration, at-
mospheric oxygen levels, and the origin-disappearance of
MIF-S.
The larger range of
34
S observed in the Proterozoic sug-
gests sulfate concentrations higher than 200 M (Habicht et
al., 2002). Sulfate was sustained at levels sufficient to allow
the widespread expression of a biological signal (14 mM),
but not at levels as high as those in the Phanerozoic (~530
mM: Horita et al., 2002; Shen et al., 2002, 2003; Canfield,
2004; Kah et al., 2004; Gellatly and Lyons, 2005; Johnston et
al., 2006). In addition, sulfate concentrations varied within
the oceans over Proterozoic time (Logan et al., 1995; Hurtgen
et al., 2002, 2005; Shen et al., 2002, 2003; Johnston et al. 2006;
Halverson and Hurtgen, 2007), and these variable sulfate con-
centrations reflected drawdown in shallow or restricted basins
514 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 514
primitive mantle
(
33
S = 0)
re
c
y
c
le
d
s
e
d
im
e
n
ts
(p
o
s
itiv
e
?

33S
)
r
e
c
y
c
le
d

a
lt
e
r
e
d

c
r
u
s
t
(n
e
g
a
t
iv
e
?

3
3
S
)
sulfate from
sulfide weathering
(positive?
33
S)
m
etam
orphism
(redistribution of
33
S)
atmospheric sulfur chemistry
Production of MIF
(negative
33
S) (positive
33
S)
geographic
variability of
33
S
volcanic SO
2
and H
2
S
(
33
S = 0?)
Proportion of atmospheric sulfur increases
as a result of proximity to sulfur sources
or to changes in sulfur flux
(low frequency variations
sedimentary Py from S
0
(positive
33
S)
Sulfate reduction
dilutes
33
S signal
Sulfate reduction
dilutes
33
S signal
Atmosphere: atmospheric deposition
of sulfur with
33
S
flux depends on proximity
to global and regional sources

33
S signal may vary
solar
radiation
atmospheric sources
proxim
ity to
Oceanic sulfate
(near zero or negative
33
S)
Oceanic sulfate
(near zero or negative
33
S)
Oceanic sulfide
(positive
33
S)
Oceanic sulfide
(positive
33
S)
Hydrothermal
ore deposits
hydrothermal circulation
(trapping of
33
S signals from sulfur pools)
Ocean: Two sulfur pools
1. Sulfate (<0.2 mM)
2. Water column sulfate reduction
produces FeSH (if ferruginous)
or HS (if locally euxinic)
(long-term signal reflects proportion of
sulfur from atmosphere compared to
proportion of sulfur from sulfate reduction,
short term signal reflects water column
and sediment related processing of
sulfur via sulfate reduction and
incorporation of atmospheric sulfur into
pyrite.)
Ocean: Two sulfur pools
1. Sulfate (<0.2 mM)
2. Water column sulfate reduction
produces FeSH
+
(if ferruginous)
or HS
-
(if locally euxinic)
(long-term signal reflects proportion of
sulfur from atmosphere compared to
proportion of sulfur from sulfate reduction,
short term signal reflects water column
and sediment related processing of
sulfur via sulfate reduction and
incorporation of atmospheric sulfur into
pyrite.)
FIG. 5. Model that describes a possible Archean sulfur cycle. The MIF signal is of atmospheric origin and the transfer of
this signal to the sediments includes components related to atmospheric, water column, and sedimentary processes. Larger
scale (longer-term) trends for
33
S are interpreted to reflect long-term variations in delivery of sulfur from the atmosphere
reflecting global and regional controls as well as reaction pathways. Smaller scale (shorter-term) variability for
33
S is inter-
preted to reflect changes in water column and sedimentary pathways for production and preservation of pyrite.
where sulfate reduction rates were high and resupply by cir-
culation was low. The average
34
S of sedimentary pyrite for
Paleoproterzoic rocks remains similar to the value seen in the
Archean (i.e., 2), implying abundant pyrite burial. The av-
erage
34
S of sedimentary pyrite for Mesoproterozoic and
Neoproterzoic rocks increased to higher values (~4-7; Fig.
1B), which implies that both sulfide and sulfate are
34
S en-
riched relative to the long-term sulfur inputs. This indicates
that a pool of
34
S-depleted sulfur is missing from the known
sedimentary pyrite record. Such a pool may have been an off-
shore pool that was lost to deeper water, sedimentary facies,
and/or sulfide deposits. It would imply that sulfur isotopic
compositions and concentrations were heterogeneous in the
Proterozoic oceans. An interesting suggestion made by Can-
field (2004) is that the size of the exogenic sulfur reservoir
was affected by loss of a deep-water sulfide pool to rocks that
were ultimately subducted.
The record of four isotopes since ~600 Ma supports a sta-
bilization in isotopic variability for seawater sulfate that is re-
lated to generally higher (430 mM) oceanic sulfate concen-
trations and an average
34
S that declines to negative values
(Fig. 1B) for the first time, implying a less significant role for
pyrite burial. This interpretation is consistent with other di-
rect types of evidence for higher sulfate levels that come from
fluid inclusions (e.g., Horita et al., 1991, 1996, 2002; Lowen-
stein et al., 2005) and suggests high sulfate concentrations at
least as early as the latest Neoproterozoic. (Horita et al., 2002;
Lowenstein et al., 2005).
It has been suggested that the rise in sulfate concentration
since the latest Neoproterozoic and the change in the fraction
of sulfur lost to pyrite burial are related to oxygenation of
shelf environments and the establishment of conditions fa-
vorable for animal radiation and bioturbation (Canfield and
Farquhar, 2009). Furthermore, changes in the magnitude of
the fractionation associated with buried pyrite in the
Phanerozoic are related to changes in oxidation pathways,
which may have similar links to the evolution of open diage-
netic systems on a global scale (e.g., Wu et al., 2010). The
change in the magnitude of the fractionation between sulfate
and buried pyrite at ~540 Ma could be related to bioturbation
because of the way that it affects recycling of sulfur from sul-
fate reduction (Fig. 6), and also because of links between bio-
turbation, sulfate concentration, and the magnitude of frac-
tionations produced by sulfate-reducing bacteria (SRB)
(Canfield and Farquhar, 2009). This presents an interesting
question related to the observed change in the range of
34
S
in the Neoproterozoic (e.g., ~800 Ma in Canfield and Teske,
1996), the change in the fractionation between sulfate and
buried pyrite (Fig. 1), and the change in
33
S starting at ~1.4
Ga reflective of a change in oxidation pathways and dispro-
portionation (Johnston et al., 2005a). If we accept that the
change in
33
S reflects the onset of disproportionation and
the change in the fractionation between sulfate and buried
pyrite is associated with bioturbation and the rise of oceanic
sulfate, then we are left with the question of what is the sig-
nificance of the change in the range of
34
S at ~800 Ma. Can-
field and Teske (1996) suggested that this change is the ex-
pression of disproportionation in the sedimentary record.
This may be the case if the signal seen by Johnston et al.
(2005a) was for water column disproportionation, where the
sulfide pool was large and relatively homogenous. This would
imply a retreat of the sulfur ecology, with disproportionation,
into sedimentary diagenetic environments at ~800 Ma.
The Record of Sulfur Cycle Evolution
from Sulfide-Bearing Ore Deposits
Several types of sulfide-bearing ore deposits can be linked
to oceanic sulfur chemistry and/or to the isotope variations
observed in the surface sulfur cycle. These include vol-
canogenic massive sulfide deposits, clastic-dominated Pb-Zn
deposits, sediment-hosted copper deposits, and Mississippi
Valley-type (MVT) deposits.
Volcanogenic massive sulfide (VMS) deposits
Volcanogenic massive sulfide (VMS) deposits occur through-
out geologic history, with examples identified in the Archean,
Proterozoic, and Phanerozoic. The VMS deposits are associ-
ated with volocanogenic settings and hosted in volcano sedi-
mentary piles (see Huston et al., 2010). These deposits are
metal-sulfide deposits, principally Fe, Cu, Pb, and Zn, associ-
ated with large-scale hydrothermal circulation systems that
involve seawater and may involve sulfur leached from igneous
and other crustal rocks or from reduction of seawater sulfate.
Ono et al. (2007) show for sulfide in modern black smokers
from the East Pacific Rise, that 73 to 89 percent of the sulfide
originates from leaching of juvenile sources and 11 to 27 per-
cent originates from reduction of seawater sulfate. Peters et
al. (2010) apply similar techniques to the Mid-Atlantic Ridge
and find slightly higher proportions of sulfide derived from
the reduction of seawater sulfate that they attribute to the
deeper circulation of the hydrothermal system. These sys-
tems are thought to be present-day analogs of ancient VMS
systems.
Present-day hydrothermal systems in active extensional set-
tings are also thought to be analogs for ancient VMS deposit
formation. These deposits generally grade from sulfide stock-
work zones into overlying strata-bound massive sulfide and
exhalites. Sulfide minerals may form in the water column as a
result of a variety of processes, including mixing or cooling of
metal- and sulfide-bearing fluids, and may depend on factors
such as the metal/sulfur ratio and the presence of sulfidic bot-
tom waters (e.g., Goodfellow and Peter, 1996; Tornos et al.,
1998). The recharge zones where seawater infiltrates the sys-
tem may also be sites for precipitation of sulfate minerals,
such as anhydrite that precipitates by retrograde solubility
along flow paths that progress up-temperature. Much of the
anhydrite is not preserved in older deposits because it is lost
once these systems cool. Other sulfate minerals such as barite
form when solutions containing sulfate mix with barium-rich
solutions. The isotopic composition of sulfur in these deposits
records information about the state of the oceanic sulfur
cycle.
SEDEX CD-Pb-Zn sulfide deposits
Sedimentary exhalative (SEDEX) deposits are a type of
sedimentary rock-hosted sulfide-bearing deposit that forms
when metal and sulfide-bearing fluids are mixed, cool, or
evolve to a different pH in syngenetic to diagenetic sedimen-
tary environments. The fluids that carry metals have migrated
predominantly from underlying clastic sedimentary rocks and
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 515
0361-0128/98/000/000-00 $6.00 515
fluid flow is driven by high geothermal gradients and focused
along syndepositional extensional faults (Sangster, 2002; Yang
et al., 2004). Formation of ore minerals may occur below the
sea floor in permeable zones, but some SEDEX deposition is
associated with discharge of fluids into seawater to yield a
strata-bound deposit. Synsedimentary features are preserved
in some SEDEX deposits, indicating a connection between
the sulfide minerals and the sediments that make up these se-
quences. The connection with a vent is not always clear and
evidence suggests that, in some cases, the brines migrate
516 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 516
sulfate input
evaporite sink

Seawater sulfate
Porewater sulfate
Porewater sulfide
Buried Pyrite
intermediates
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0 10 20 30 40 50 60 70 80

34
S
sulfate

3
3
S
m
o
r
e

c
lo
s
e
d

s
y
s
t
e
m

M
o
r
e

o
p
e
n

s
y
s
t
e
m

-

l
a
r
g
e
r

3
4
S
p
y
r
i
t
e

b
u
r
i
a
l

M
o
r
e

c
l
o
s
e
d

s
y
s
t
e
m

-

s
m
a
l
l
e
r

3
4
S
p
y
r
i
t
e

b
u
r
i
a
l
With bioturbation Without bioturbation
Seawater sulfate
Seawater sulfate
Porewater sulfate
Porewater sulfate
Porewater sulfide Porewater sulfide
Buried Pyrite
Buried Pyrite

3
4
S
p
y
r
i
t
e

b
u
r
i
a
l
0
none
1
all
Fraction of return flow to
seawater sulfate
u
n
b
i
o
t
u
r
b
a
t
e
d

s
e
d
i
m
e
n
t
b
i
o
t
u
r
b
a
t
e
d

s
e
d
i
m
e
n
t
larger
smaller
A B
C D
FIG. 6. A model system (box model representation) to illustrate how bioturbation may change the degree of openness of
a diagenetic setting and alter the magnitude of the fractionation preserved between buried pyrite and seawater sulfate. B.
Schematic plot showing relationship of fractionation between sulfate and pyrite and fraction of sulfur recycled to the oceanic
pool. Another direct influence on fractionation related to bioturbation is coupled to the possibility of changes in sulfate con-
centration that feed back to the dependence of isotope discrimination by sulfate-reducing bacteria on sulfate concentrations
(e.g., Canfield and Farquhar, 2009). C. Similar model system that incorporates disproportionation and D. Calculated fields
for overlying sulfate pool on plot of
33
S versus
34
S. Calculated using fractionations for sulfate reduction from Farquhar and
Johnston (2008) from experiments with natural populations of sulfate reducers and disproportionation data from Johnston et
al. (2005b). Dark line is for sulfate reduction only. Dashed line includes reoxidation and disproportionation. Other fraction-
ations yield different size fields, but the general shapes are preserved.
considerable distances from the discharge site without mixing
with seawater (Sangster, 2002).
Leach et al. (2010) argue that the term clastic-dominated
lead-zinc (CD Pb-Zn) deposits is more appropriate because it
includes SEDEX deposits, but provides a descriptive classifi-
cation that does not imply a genetic link to exhalative processes.
They note that many of the deposits lumped with SEDEX de-
posits in other studies lack clear evidence for an exhalative ori-
gin, despite many forming in the sediments (Leach et al.,
2005). They suggest a further qualification that uses a desig-
nation to describe CD Pb-Zn deposits in terms of the tectonic
setting in which they form, such as passive margin, continen-
tal rift, continental sag basin, and back-arc basins. In this
paper, we will use the Leach et al. (2010) terminology.
MVT sulfide deposits
MVT deposits are base metal (Fe, Zn, Pb) sulfide deposits
hosted in carbonate-dominated sedimentary rocks. The MVT
deposits are epigenetic deposits that are generally strata
bound, but commonly not stratiform, with sulfide minerals
replacing carbonates and filling open spaces in the host rocks
(Sangster, 1990). Most MVT deposits are thought to form as
a result of the mixing between sulfide-bearing and metal-
bearing basinal brines, much like modern oil field brines that
are mobilized by tectonic events (Leach et al., 2001, 2005;
Paradis and Nelson, 2007). Leach et al. (2010) suggest that
sulfate may be stored in these basins until remobilized by a
tectonic trigger, and the appearance of MVT deposits in the
Phanerozoic implies a link to a threshold for establishing a
pool of stored sulfate. The ultimate source of sulfide is
thought to be seawater sulfate because of its isotopic compo-
sition and, in most cases, is thought to be the product of bio-
logical or thermochemical sulfate reduction. The epigenetic
processes by which most MVT deposits are thought to form
make them inherently difficult to date and also remove their
chemistry several steps from ocean chemistry and the surface
sulfur cycle, but links to the sulfur cycle remain and are ex-
plored forthwith.
The MVT deposits are not recognized before ~2.2 Ga (the
Pering and Bushy Park deposits) and peak at times when
ocean sulfate concentrations were highest (Kesler and Reich,
2006). Leach et al. (2001) have shown that the best dated of
the Phanerozoic deposits cluster during periods of tectono -
thermally related flow of brines almost exclusively associated
with closure of ocean basins. Two principal episodes of
Phanerozoic MVT formation can be defined at ~400 to 300
Ma and ~120 to 60 Ma. These two episodes correspond to pe-
riods when seawater sulfate concentrations (e.g., Horita et al.,
2002; Lowenstein et al., 2003) were high, but not at their
maxima (Fig. 7). They also correspond with times of tecton-
ism and the link between tectonism and basinal brine migra-
tion is the primary reason for their formation, although suffi-
cient sulfate is also required (Kesler et al., 1995; Leach et al.,
2001, 2005, 2010). The source of these basinal brines varies
from deposit to deposit. It has been inferred on the basis of
fluid inclusion Cl/Br ratios that the dominant source of some
of these brines is evaporated seawater, rather than the disso-
lution of evaporites (Kesler et al., 1995; Leach et al., 2001). In
other cases, there appears to be evidence for brines that were
derived from the dissolution of evaporites. In the case of the
most ancient MVT deposits, at Bushy Park and Pering in
South Africa, the mineralization appears to have occurred at
deeper crustal levels (2.84.8 km) than at other MVT deposits
(Huizenga et al., 2006) and the
34
S,
33
S, and
36
S of the sul-
fur (Schaefer; 2002; Kim et al., 2009) appear to be more
closely related to sulfur from host rocks than from contempo-
raneous or older seawater sulfate sources that are inferred to
have zero or negative
33
S, respectively.
Sediment-hosted copper deposits
The distribution of sediment-hosted copper deposits (Hitz-
man et al., 2010) is also associated with the evolution of the
sulfur and oxygen cycles. These are stratiform deposits of
copper sulfides in siliciclastic and/or dolomitic sedimentary
rocks. They first appear in the Paleoproterozoic after the rise
of oxygen and sulfate, and have formed through the present
day (Brown, 1997). Their formation is attributed to the reac-
tion of oxidized solutions containing copper, which are gener-
ated during diagenetic reddening reactions of sediments
(redbeds), with sulfide-containing solutions produced at, or
introduced to, the site of deposition (Hitzman et al., 2005;
Brown, 2005, 2009). Sulfur sources for the sulfide-bearing so-
lutions include evaporites, basinal brines, and sulfidic hydro-
carbons. Connections between sediment-hosted copper de-
posits and the oceanic sulfur cycle are indirect because these
deposits typically form in continental rift settings and are
thought to be more common when rift settings are at low lat-
itude, allowing evaporative concentration of sulfate (Brown,
1997). This implies a connection to sulfate concentration, but
also calls for a concentrating mechanism. For example, sulfur
isotope studies of the Central African Copperbelt (Cailteux et
al., 2005) suggest reduction of seawater sulfate in rift basin
settings that are at times isolated from the oceans and provide
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 517
0361-0128/98/000/000-00 $6.00 517
0
1
2
3
4
5
6
7
8
0 100 200 300 400 500 600
0
10
20
30
0 100 200 300 400 500 600
Age (Ma)
C
o
u
n
t
S
u
l
f
a
t
e

(
m
M
o
l
a
r
)
assembly of Pangea alpine -
Pangea laramide
(assimilation) assimilation
FIG. 7. Comparison between histogram of frequency of MVT mineraliza-
tion events (using data in Leach et al., 2001) and models of the evolution of
seawater sulfate concentrations from Lowenstein et al. (2005) (thick black
line) and Horita et al. (2002) (thick gray line with estimates of uncertainty
given by thin gray lines). The figure shows a correspondence between the
frequency of MVT deposit forming events and tectonic processes, rather
than a direct relationship with sulfate concentrations. It is not clear whether
sulfate lags mineralization events or vice versa.
stratigraphic evidence of control on isotopic composition by
communication between basin sulfate and oceanic sulfate.
Connections between Seawater Sulfate and
Sulfide Ore Deposits
The shared connection through hydrothermal circulation
systems to seawater for both VMS and CD Pb-Zn deposits
provides a link to seawater sulfate. The connections between
the brines that form MVT deposits, as well as the sediment-
hosted copper deposits, and preexisting evaporites and/or sea-
water sources provide a link between these deposits and the
sulfur cycle, but as we have noted above, their epigenetic ori-
gin adds complexity to the interpretation. The similarity in the
isotopic compositions of sulfide and sulfate for VMS, CD Pb-
Zn, and MVT deposits relative to that for data compiled from
the sedimentary record is illustrated in Figure 8A. Note that
pyrite may have a much broader range in values relative to ore
sulfides such as sphalerite and galena (cf. Kelley et al., 2004;
Leach et al., 2005). The common features of these two records
include the small range of
34
S for sulfide and sulfate in sam-
ples of Archean age, an expanded range of
34
S in Proterozoic
age samples that is biased to positive
34
S values, and a further
expanded range of
34
S in Phanerozoic age samples. This cor-
respondence illustrates a broad link that exists between the
sulfur in these sulfide-bearing ore systems and sulfur from
other parts of the exogenic sulfur cycle. However, the parallel
is not perfect. Whereas ore sulfides broadly overlap with the
total range of observed sedimentary pyrite and sulfate, ore
sulfides are generally more enriched in
34
S (higher
34
S) than
sedimentary pyrite. This reflects a stronger link to sulfate
rather than sulfide in surficial environments. The nature of the
connections is different for different classes of ore deposits
and also for the different parts of the ore deposits, such as
stockworks compared to exhalative lenses in VMS deposits.
The processing of sulfur in the ore depositsfor example, sul-
fate in barite capsreflects Rayleigh fractionation that con-
tributes to the variability. Changes in the distribution of
34
S
for the three types of sulfide deposits can be noted over Earth
history (Figs. 9, 10). The characteristics of these distributions
differ from those of the sedimentary record insofar as the
mean values are generally positive for both sulfate and sulfide
for all deposit types, which reflects the role played by sulfate
and leaching of igneous sulfur, ultimately of juvenile origin, as
sources of sulfur rather than of pyrite produced from reduced
sulfur contributed by sulfate reducers. As noted above, iso-
topic studies of modern hydrothermal systems at the East Pa-
cific Rise and Mid-Atlantic Ridge (Ono et al., 2007; Peters et
al., 2010) suggest that the fraction of sulfur derived from sul-
fate is less than ~40 percent of the total sulfur and the re-
mainder is derived from leaching of igneous sulfides. The dif-
ference in
34
S for sulfate and sulfide when they coexist in
these different types of deposits reflects a combination of mix-
ing and fractionation processes in the ore-forming processes.
The VMS deposits form at the highest temperatures (usually
>300C), whereas the CD Pb-Zn and MVT deposits typically
form at lower temperatures (<250C). There is also a clear
progression to more positive and more variable
34
S for sul-
fide from younger VMS deposits, reflecting inclusion of the
Bathurst deposit and deposits in northern New Brunswick,
that is attributable to a combination of higher
34
S for marine
518 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 518

3
4
S
Age (Ma)
A.
-80
-60
-40
-20
0
20
40
60
80
100
0 500 1000 1500 2000 2500 3000 3500 4000

3
4
S
Age (Ma)
B.
0
10
20
30
40
50
60
0 500 1000 1500 2000 2500 3000 3500 4000
FIG. 8. A. Compilation of
34
S vs time for base metal sulfides filled gray
circle symbols and for sulfate (barite, anhydrite, gypsum) unfilled diamond
symbols. Unfilled gray circle symbols are same as those plotted in Figure 1
for sedimentary pyrites. Black line is seawater sulfate curve from Figure 1. B.
Expanded plot with data for sulfate from ore deposits and seawater sulfate
curve presented in Figure 1. Note that we do not include data for the ~2.7
Ga Hemlo deposit, because of the inference that much of this variation re-
flects later redistribution of sulfur isotopes during amphibolite facies meta-
morphism (Thode et al., 1991). (DeChow, 1966; Tupper, 1960; Buschendorf
et al., 1963; Anger et al., 1966; Ryznar et al., 1967; Sasaki and Krouse, 1969;
Solomon et al., 1969, 1988; Guha, 1971; Kajiwara, 1971; Lusk, 1972; Arnold
et al., 1977; Carr and Smith, 1977; Campbell et al., 1978, 1980; Smith et al.,
1978; Green et al., 1981; Rye and Williams, 1981; Sverjensky, 1981; Walker
et al., 1983; Willan and Coleman, 1983; Akande and Zentilli, 1984; Goodfel-
low and Jonasson, 1984; Olson, 1984; Gardner and Hutcheon, 1985; Scott et
al., 1985; Taylor and South, 1985; Deloule et al., 1986; Shanks et al., 1987;
Eastoe and Nelson, 1988; Andrew et al., 1989; Crocetti and Holland, 1989;
Kase et al., 1990; Whelan et al., 1990; Pandalai et al., 1991; Davidson and
Dixon, 1992; Eldridge et al., 1993; Nakai et al., 1993; Cagatay and Eastoe,
1995; Frietsch et al., 1995; Bechtel et al., 1996; Dixon and Davidson, 1996;
Jones et al., 1996; Perkins, 1996; Cook and Hoefs, 1997; Kesler et al., 1997;
I.K. Anderson et al., 1998; Broadbent et al., 1998; Velasco et al., 1998, 2003;
Painter et al., 1999; Ding and Jiang, 2000; Gaboury et al., 2000; Lyons et al.,
2000; Melezhik et al., 2000; St. Marie and Kesler, 2000; Sharpe and Gem-
mell, 2000; B.R. Anderson et al., 2001; Luepke and Lyons, 2001; Blakeman
et al., 2002; Wagner et al., 2002; Peevler et al., 2003; Ma et al., 2004, 2007;
Wilkinson et al., 2005; Gilg et al., 2006; Huizenga et al., 2006; Jamieson et al.,
2006; Kesler and Reich, 2006; Leach et al., 2006; Decre et al., 2008; Tornos
et al., 2008; Johnson et al., 2009).
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 519
0361-0128/98/000/000-00 $6.00 519
0
10
20
30
40
50
60
70
80
90
-50 -40 -30 -20 -10 0 10 20 30 40 50 60 70
0
5
10
15
20
25
30
35
40
45
-50 -40 -30 -20 -10 0 10 20 30 40 50 60 70
0
5
10
15
20
25
30
35
40
0
100
200
300
400
500
-50 -40 -30 -20 -10 0 10 20 30 40 50 60 70

34
S
N
u
m
b
e
r

34
S
N
u
m
b
e
r

34
S
N
u
m
b
e
r
_
X= 7.6 8.0
sulfide:
sulfate:
_
X= 28.8 7.7
sulfide:
_
X= 4.1 6.6
sulfide:
_
X= 1.6 1.4
FIG. 9. Histograms of the frequency of observations of sulfide (gray filled bars), and sulfate (hatched bars), for Phanero-
zoic (A), Proterozoic (B), and Archean (C) VMS deposit intervals in geologic time. The heights of each histogram are nor-
malized to the total number of samples in order to put them on a similar aspect ratio. Mean and standard deviation given on
plots. Data sources listed in Figure 8.
A
B
C
sulfate and to these particular sedimentary rock-hosted VMS
deposits.
The connections that are illustrated in Figure 8A are not,
however, straightforward because ore-forming processes also
influence the
34
S of the sulfide and sulfate in these deposits.
Sulfate minerals from these deposits possess isotopic compo-
sitions that generally extend to
34
S-enriched compositions and
in some cases also include
34
S-depleted sulfate (Fig. 8B).
These variations can be attributed to reservoir effects associ-
ated with removal of
32
S-enriched sulfide produced by sulfate
reduction (Rayleigh fractionation and related effects: e.g.,
Goodfellow, 1987) and the
34
S-depleted sulfate may reflect
oxidation of
32
S-enriched sulfide. Note that reservoir effects
such as Rayleigh fractionation will be expressed in the water
column: the fraction of sulfate consumed in the water column
is limited by the availability of nutrients, export of sulfides by
sedimentation, and eddy diffusion of sulfate and sulfide, so a
significant proportion of these may be associated with pore-
waters. Johnson et al. (2009) demonstrated covariance for sul-
fur and oxygen isotopes that evolve from compositions similar
to seawater sulfate to higher
34
S and generally higher
18
O.
The variation they report for
34
S is similar to that reported in
other studies (e.g., Fig. 10) and they attribute this variation to
the processes associated with sulfate reduction in the region
where barite precipitates. Broadbent et al. (1998) describe a
process at the Century CD Pb-Zn deposits of the McArthur
Basin that involves an increase in the
34
S/
32
S of approximately
20 per mil with the evolution of the deposit. They interpret
this to reflect drawdown of the sulfate sulfur pool as a result
of sulfide sinks outcompeting sulfate resupply. This observa-
tion also may be related to the observation of high
34
S in Pro-
terozoic CD Pb-Zn deposits (e.g., Lyons et al., 2004). This
suggestion implies that the fractional drawdown of sulfate in
these systems is greater when sulfate concentrations are
lower. Although this link has not been documented (but see
Johnson and Emsbo, 2005), it would presumably draw on an
inferred correlation between concentration and a function
that relates the amount of sulfur given by the deposit size,
1/(fractional drawdown of sulfate), and 1/(fluid flux). Johnson
and Emsbo (2005) suggest that the drawdown of basinal or
global oceanic sulfate would be minimal, implying that distil-
lation occurs in the sediments and not in the water column.
The persistence of high
34
S in Phanerozoic CD Pb-Zn de-
posits (Fig. 10) suggests that the record preserved by these
deposits may not uniquely support this suggestion, although
they would not disprove it, either.
Insights from sulfide ore deposits for the Archean S-cycle
There are some features of VMS deposits that appear to
provide important information about the state of the oceanic
sulfur cycle. Some of the oldest (34003200 Ma) VMS de-
posits are preserved in Western Australia; the Kangaroo
520 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 520
0
10
20
30
40
50
60
70
80
-50 -40 -30 -20 -10 0 10 20 30 40 50 60 70
-50 -40 -30 -20 -10 0 10 20 30 40 50 60 70
0
50
100
150
200
250
300
350
-50 -40 -30 -20 -10 0 10 20 30 40 50 60 70
0
5
10
15
20
0
100
200
300
400
500
-50 -40 -30 -20 -10 0 10 20 30 40 50 60 70
0
10
20
30
40
50
0
50
100
150
200
250
300
350
400
450
-50 -40 -30 -20 -10 0 10 20 30 40 50 60 70
0
20
40
60
80
100
120

34
S
N
u
m
b
e
r

34
S

34
S

34
S
N
u
m
b
e
r
N
u
m
b
e
r
N
u
m
b
e
r
A. MVT (Phanerozoic)
B. MVT ((Proterozoic)
C. CD-Pb-Zn (Phanerozoic)
D. CD-Pb-Zn (Proterozoic)
_
X= 23.6 10.4
_
X= 7.9 9.9 sulfide:
sulfate:
_
X= 12.2 10.7
_
X= 25.1 7.2
sulfide:
sulfate:
_
X= 6.4 16.4 sulfide:
sulfate:
_
X= 25.3 10.3
_
X= 23.7 0.5
sulfide:
sulfate:
_
X= 15.3 11.3
FIG. 10. Histograms of the frequency of observations of sulfide (gray filled bars), and sulfate (hatched bars), for Phanero-
zoic MVT (A), Proterozoic MVT (B), Phanerozoic CD-Pb-Zn (C), and Proterozoic CD-Pb-Zn (D) deposits. The heights of
each histogram are normalized to the total number of samples in order to put them on a similar aspect ratio. Mean and stan-
dard deviation given on plots. Data sources listed in Figure 8.
Caves and the Sulphur Springs deposits (Vearncombe et al.,
1995; Golding and Young, 2005). They are characterized by
mineralogical and textural features that are similar to younger
VMS deposits with Cu, Pb, and Zn sulfide mineralization and
appreciable barite. The deposits formed in a back-arc setting
at a water depth greater than ~1,000 m and at or just below
the sea floor (Vearncombe et al., 1995; Buick et al., 2002).
They are contemporaneous with sulfate-bearing shallow-
water deposits from other localities in Western Australia and
South Africa (Heinrichs and Reimer, 1977; Buick and Dun-
lop, 1990).
The presence of barite in the Sulphur Springs and Kangaroo
Caves deposits indicates that barium was mobilized in the
subsea-floor hydrothermal system and that sufficient sulfate
was present at the site of deposition to precipitate barite
(Vearncombe et al., 1995; Huston et al., 2002). It can be ar-
gued on the basis of
33
S that the source of sulfate is marine
(Farquhar et al., 2000; Golding et al., 2006), which is consis-
tent with suggestions from other evidence (Vearncombe et al.,
1995; Huston et al., 2002). These arguments also suggest that
the early oceans may have had high sulfate concentrations and
that this requires an oxygenated early Earth. We do not sub-
scribe to this suggestion because an oxygenated atmosphere is
not required for production of at least some sulfate in the ex-
ogenic cycle (e.g., Walker and Brimblecombe, 1985; Kasting,
2001; Farquhar et al., 2000, 2001). Sulfate will be produced
by photochemical pathways from volcanic emissions even on
an anoxic early Earth (Vanderwood and Thiemens, 1980;
Walker and Brimblecombe, 1985; Farquhar et al., 2001; Kast-
ing, 2001), and evidence for these reactions is present in the

33
S of the sulfate in these deposits and elsewhere (Hoering,
1989; Bao et al., 2007). The precipitation of sulfate requires
only that the solubility product for barite be exceeded.
Whereas the amounts of sulfate present in these deposits
requires a sufficient supply of sulfate, the presence of barite
may not imply oceanic sulfate concentrations >200 M. Hus-
ton and Logan (2004) propose that the oceans may have been
barium enriched and sulfate limited. These authors present a
number of arguments, including many listed above, to make
the point that evidence does not exist to support sulfate-rich
oceans at this time. It is valuable to realize, however, that the
point at which the solubility product of BaSO
4
is balanced be-
tween Ba
2+
and SO
4
2-
is in the range of tens of M, which is
below, but not significantly, the solubility of Ba
2+
predicted
for an ocean equilibrated with 0.1 bar of CO
2
. One-tenth of
a bar of carbon dioxide is a level contained in Archean atmos-
pheric energy balance models (e.g., Domagol-Goldman et al.,
2008; Haqq-Misra et al., 2008). This relationship between
CO
2
, Ba
2+
, and SO
4
2
suggests that both species coexisted at
levels of a few tens to a few hundred M and that barite pre-
cipitation occurred when these levels were exceeded.
Two other observations bear on the nature of the oceanic
sulfate pool at this time in early Earth history. First, barite
samples from other Paleoarchean localities in Australia,
South Africa, and India (Hoering, 1989; Farquhar et al., 2000;
Bao et al., 2007; Ueno et al., 2008, Shen et al., 2009) also
carry a
33
S signature that implies the sulfate was derived
from seawater. In particular, the
33
S signal is consistently
negative and has been inferred to represent the seawater sul-
fate pool. The
34
S of this sulfate is not strongly positive
(~38) and thus the isotopic composition of seawater sul-
fate was not strongly fractionated by global-scale burial of
pyrite. Second, Foriel et al. (2004) use the Cl/Br of fluid in-
clusions from the quartz-carbonate pods of the 3.49 Ga
Dresser Formation, a shallow-water facies that is slightly
older than the Kangaroo Caves and Sulphur Springs deposi-
tional environments, to argue for mixing of Ba-rich hy-
drothermal fluids with low sulfate seawater that had been
subject to evaporative concentration, consistent with other
constraints on the nature of the local depositional environ-
ment. The fluid inclusion data support an upper limit of 8
mM for seawater sulfide. This is not as low as the estimates
made using other lines of evidence; it is also the upper detec-
tion limit of the analytical technique.
A number of Neoarchean VMS deposits do not appear to
contain similar barite mineralization. This reflects lower sul-
fate concentrations or the nature of the hydrothermal circu-
lation systems such that the size of the events in the
Neoarchean may have been unusually large and that they may
have influenced the chemistry of the overlying water, thus
making barite precipitation unlikely. Both of these sugges-
tions are difficult to test, and each may carry consequences
for the sulfur cycle. For the first suggestion of lower global
seawater sulfate concentrations, the reasoning is that during
the Neoarchean there was a return to more anoxic conditions
from an earlier Archean, more oxygenated environment. As
described above, we do not believe that the evidence for an
oxygenated Paleoarchean is compelling and another reason
for a change in sulfate concentrations would be required. It is
possible, but not proven, that oceanic sulfate concentrations
may have declined in the Neoarchean because of the estab-
lishment of more effective sinks for sulfate, perhaps associ-
ated with the establishment of a niche for sulfate reducers.
This is difficult to test, but evidence for the hypothesis should
be present in the
34
S of seawater sulfate. Some isotopic data
for pyrite in carbonates, sulfide from marine sediments, and
sulfides and rare barite in VMS deposits suggest that the
34
S
of Neoarchean sulfate had higher values (~1015; Thode
et al., 1991; Ono et al., 2003; Farquhar and Wing, 2005; Kauf-
man et al., 2007). This may signal the onset of bacterial sul-
fate reduction in the oceanic sulfur cycle and would imply a
growing sink of pyrite burial and larger fractionations associ-
ated with this sink.
Cameron and Hattori (1987) report
34
S analyses of pyrite
and sulfate (barite and anhydrite) from Hemlo and related
gold-rich possible VMS deposits in the Superior province that
provide an exception to these more general observations of a
lack of sulfate and a lack of significant sulfur isotope fraction-
ation in Neoarchean deposits. They interpret these deposits
to be the result of metal deposition in an environment with lo-
cally high sulfate concentrations that indicate an Archean
ocean with isolated sulfate-rich basins, despite a relatively sul-
fate poor overall concentration. The range of variability for

34
S approaches 30 per mil and is interpreted to be a result of
abiological processes. Thode et al. (1991) note that the frac-
tionation of
34
S/
32
S between coexisting barite and pyrite is uni-
form, despite compositions that vary from sample to sample.
They also note the observed fractionations are remarkably
similar to the predicted fractionation on the basis of equi-
librium isotope exchange at postdepositional metamorphic
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 521
0361-0128/98/000/000-00 $6.00 521
temperatures at Hemlo and that the compositions correlate
with mass balance in a way suggestive of local reequilibration.
A further future test of this hypothesis could be provided
using an approach similar to that of Jamieson et al. (2006),
who relied on a combination of
33
S and
34
S to evaluate the
approach to a single-step equilibrium isotope effect. Never-
theless, Thode et al. (1991) describe a relationship between
the isotopic composition of barite and pyrite versus modal
pyrite-barite that is interpreted in terms of an initial barite
composition (~10). In summary, Hemlo provides a rare
window into the depositional environment of a deposit that
has been subjected to a rather complex metamorphic history
(e.g., Muir, 2002; Heiligman et al., 2008).
Insights from sulfide ore deposits for the Proterozoic S-cycle
Information provided by Proterozoic VMS deposits is of a
different type than provided by Archean VMS deposits. Rel-
atively fewer VMS deposits are observed in Proterozoic rocks,
but this is not inferred to reflect a change in the oceans or in
the exogenic element cycles. Rather, it is inferred to reflect
changes in the geotectonic conditions conducive to the for-
mation and preservation of VMS deposits. Where they occur,
Proterozoic VMS are as informative about the nature of the
oceanic environments as are Archean VMS, and they provide
different information than that from CD Pb-Zn deposits and
sedimentary geochemistry. For example, Slack et al. (2007)
have argued that a lack of Ce anomalies and the presence of
hematite in jasper associated with the deep-water (>1,000 m),
late Paleoproterozoic (~1740 Ma) Jerome VMS deposit, Ari-
zona, indicate the oceans were locally oxygenated in the sta-
bility field of hematite and did not support the same global
Mn oxide sink for Ce that is present today. They also infer, on
the basis of seven other similar deposits ranging in age from
1792 to 1241 Ma, that these observations can be extrapolated
over a much broader time interval. Using these constraints
and log(
O2
)-pH (Eh-pH) diagrams (Fig. 11), it is inferred
that oxygen in the Paleoproterozoic is very low (anoxic), but
still variable over many orders of magnitude. The constraints
on the presence of oxygen, therefore, do not appear to be re-
lated to the stability of hematite, but to the mass balance con-
straints for hematite formation. The absence of free sulfide in
water overlying Jerome and similar deposits has been argued
on the basis of the absence of pyrite (Slack et al., 2007). How-
ever, in at least one present-day, iron-rich setting (Lake
Matano, Indonesia), iron (hydr)oxides have been shown to co-
exist with very low levels of free sulfide (<0.1 M), although
there is also strong evidence for the presence of an FeS
species (Crowe et al., 2008). In another study, Slack and Can-
non (2009) make the point that the character of exhalatives
from VMS systems displays a fundamental change starting at
the time of the Sudbury impact at ca. 1850 Ma. This change
to more oxidized iron, rather than sulfidized iron, is also pre-
sented as evidence for nonsulfidic conditions at the sites
where these exhalites are formed and conversely for condi-
tions with sulfide plus iron associated with the formation of
older exhalites. Slack and Cannon (2009) suggest that the
Sudbury impact event may have been a trigger for the change
in ocean chemistry.
The CD Pb-Zn deposits also appear to have links to pore
water and water column sulfate and sulfide. The CD Pb-Zn
deposits first appear during the late Paleoproterozoic, with a
number of Zn- and Pb-rich sulfide deposits located in the
McArthur basin of north-central Australia and are present
thereafter in a number of Proterozoic and Phanerozoic settings
522 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 522
0 2 4 6 8 10 12 14
pH
Mn
++
Alabandite
Bixbyite
Hausmannite
Mn(OH)
2
(am)
Pyrolusite
Rhodochrosite
25C
0 2 4 6 8 10 12
100
80
60
40
20
0
pH
l
o
g

a

O
2
(
a
q
)
FeCl
+
FeCl
2
+
Ferrite-Mg
Goethite
Magnetite
Pyrite
25C
2 4 6 8 10 12 14
pH
25C
2 4 6 8 10 12 14
25C
S
t
a
b
ilit
y
fie
ld
fo
r
G
o
e
t
h
it
e
/
F
e
r
r
it
e
-
M
g
, M
a
g
n
e
t
it
e
,
w
it
h
o
u
t
M
n
o
x
y
-
h
y
d
r
o
x
id
e
s
100
80
60
40
20
0
A B C
FIG. 11. Plots of oxygen activity in a seawater model system for Fe-O2-Mn-H2O as a function of pH at 25C, showing the
stability fields of Fe phases (A) Mn phases (B), and overlap (C) for goethite (-FeOOH) with pKsp = 44.1, where amorphous
Mn(OH)2 are not stable. Assuming total dissolved sulfate equals 1mM 10 mM at aqueous-solid boundaries. Magnetite
(Fe3O4) and pyrite (FeS2) are metasable, implying that other metastable Fe-S and Mn-S species can occur chiefly in sulfate
seawater complexes. The solubility calculation of goethite (-FeOOH) in seawater composition was estimated using the
Stumm and Morgan (1981) equation: Log Ksp = log Ksp + s (S/2.303RT), where S is the surface area of mole of fine-grained
to well crystalline goethite material (m
2
/mol), s is an assumed particle size correction term (cal/m
2
), T is temperature, and
R is 1.9872 cal/molK. Precipitation of goethite and metastable Fe-O and Fe-S mineral is due to amphoteric behavior of oxy-
hydroxides. In seawater solutions high in chloride and sulfate, initial precipitation ferrihydrite/hydrous ferric oxide HFO
Fe(OH)3 from rapid hydrolysis of Fe(III) or oxidation of Fe(II) leads to the formation of goethite at cold temperature and
hematite at high-temperature conditions. At 1 mM sulfate concentrations, chloride anions substitute for hydroxyl radicals,
forming metastable solids such as Fe(OH)(Cl) and FeCl3 complexes at low pH conditions.
(e.g., Sangster, 1990; Lott, 1999; Cooke et al., 2000; Lyons et
al., 2004). Leach et al. (2010) note that CD deposits in conti-
nental rift and continental sag basins are only present in the
Paleoproterozoic and Mesoproterozoic, whereas passive mar-
gin-type CD Pb-Zn deposits are hosted in rocks from the Pa-
leoproterozoic to the Phanerozoic. Models for the formation
of these deposits invoke mixing of sulfidic and metalliferous
solutions, in many cases with links to the presence of sulfide,
either in the sediments or in the overlying seawater (Good-
fellow and Jonasson, 1984; Goodfellow, 1987; Shanks et al.,
1987; Goodfellow et al., 1993). The sulfide from the McArthur
basin deposits has been inferred to originate either within the
sediments through sulfate reduction or from bottom waters
that may have been sulfidic (Lyons et al., 2004). Both biolog-
ical and thermochemical sulfate reduction are feasible in
these environments (Lyons et al., 2004; Huston et al., 2006).
It is interesting to consider whether the isotopic signals pro-
duced by thermochemical sulfate reduction (TSR) may pro-
duce anomalous
33
S. We know of one published analysis that
may be attributable to TSR, which is for the Urquhart Shale
of the Mt Isa inlier (PPRG 1284) with a measured
33
S within
error of zero. The distribution of CD-Pb-Zn deposits may also
be related to the presence of euxinic deep waters (e.g., Turner,
1992; Goodfellow et al., 1993; Large et al., 1998; Canet et al.,
2004; Lyons et al., 2004; Huston et al., 2006), although Slack
et al. (2004) provide an alternative interpretation for the giant
Mississippian Red Dog deposit, Alaska, which is based on the
low Mo/Ti ratio in host shales.
The chemistry of the Earths middle Proterozoic (latest Pa-
leoproterozoic, Mesoproterozoic, and early Neoproterozoic)
oceans has been a topic of a debate centered on the question
of whether they were sulfidic in their entirety, or in part (Hol-
land, 1984, 2005, 2006; Canfield, 1998). Holland (1984, 2006)
suggested that ferrous iron was removed from the deep ocean
and replaced by oxygen at ~1.9 to 1.8 Ga because of greater
oxygen availability in the aftermath of the Great Oxidation
Event (GOE). Slack and Cannon (2009) suggested that the
oceans changed to a nonsulfidic state with ~1 M dissolved
oxygen, because of the Sudbury impact, but it is not clear
what feedbacks would have stabilized the system once the
transition took place. The situation that developed at this time
would have been mainly the same as the modern, oxygenated
deep ocean, perhaps with slightly lower oxygen concentra-
tions. Canfield (1998) proposed that a ferruginous deep
ocean was not replaced by an oxygenated deep ocean during
the Proterozoic, but by a sulfidic deep ocean as a result of an
excess of sulfide over both iron and oxygen. According to
Canfield (1998), the establishment of an oxygenated deep
ocean did not occur until significantly later in geologic history,
likely closer to 600 Ma. The switch that triggered the transi-
tion from ferrous to sulfidic was a change in sulfide supply as
a result of higher sulfate delivery to the oceans. It was argued
that if phosphorus availability was similar to that of today,
then the sinks for oxygen in the deep ocean would exceed
oxygen-supplied by circulation and ventilation of the deep
oceans would not occur. Instead, anoxic oceans would result,
and sulfate delivery and reduction would result in sulfide
titrating iron and replacing it as the redox control in the deep
oceans. More recently, Holland (2006) argued that evidence
for phosphorous in Mesoproterozoic ocean sediments is
lacking and, therefore, there would not have been a large
enough flux of sinking organic matter to consume all oxygen
and to support sulfide production sufficient to establish sul-
fidic conditions.
Each of these suggestions explains some, but not all fea-
tures of the sulfur record preserved by CD Pb-Zn deposits
and oceanic sediments. Hollands (1984) suggestion does not
explain the evidence from deep sedimentary successions
(e.g., Poulton et al., 2004) or from Mo concentrations and
Mo/TOC ratios (Arnold et al., 2004; Scott et al., 2008), or ev-
idence for biomarkers from phototrophic sulfide oxidizing
bacteria, which would imply sulfide in some surface waters
(Brocks et al., 2004, 2005). It also doesnt explain the appar-
ent missing pool or
34
S-depleted sulfide evidenced by the
paucity of sedimentary sulfur with negative
34
S and a mean
value for
34
S of both sulfide and sulfate that has a long-term
enrichment relative to the long-term source of juvenile sulfur
with near zero
34
S (Logan et al., 1995; Lyons et al., 2004;
Fig, 1). Canfields (1998) suggestion does not explain some
observations from ore deposits, such as the presence of iron
oxide rather than iron sulfide in distal parts of exhalites (Slack
et al., 2007), and nonsulfide Zn deposits that appear later in
the Proterozoic (Holland, 2005).
A hybrid model of the Proterozoic ocean
A hybrid model has been proposed that attempts to recon-
cile these observations (Holland, 2005) and is rooted in the
observations from both sedimentary rocks and ore deposits.
This model invokes a combination of sulfidic and nonsulfidic
anoxic deep-water pools. There is some debate about how
these pools may have been distributed in the oceans, and also
as to how the oceans evolved following the reorganization at
~1.9 Ga. This model calls for a situation in which sulfidic and
nonsulfidic anoxic waters exist in different parts of the ocean.
Holland (2005) argues that it was not just the phosphorus
availability in the Proterozoic ocean that controlled the levels
of oxygen in downwelling waters, but also the flux of sinking
organic matter that determines the strength of the deep-
water oxygen sinks. He proposes that a slower rate of sinking
of organic matter from the surface ocean would lessen the
oxygen sink and allow for a delicate balance to be met be-
tween oxygen, sulfide, and ferrous iron that favored oxygen,
but at very low levels. Levels were low enough such that Mn
oxide was not present in high enough abundance on a global
ocean scale to scavenge elements such as Mo and Ce. This hy-
brid model that includes deep ocean sulfidic and nonsulfidic
anoxic pools is the presently accepted model for the Protero-
zoic oceans and is necessary to provide an explanation consis-
tent with the geologic and geochemical record (Table 2).
The location of the sulfidic parts of the Proterozoic ocean
would be related to sites where sulfide supply exceeded the
supply of ferrous iron and oxygen. Where the sulfate supply
was nonlimiting, the supply of sulfide would be controlled by
sulfate reduction rates and would depend on the availability
of electron donors, principally organic carbon supplied from
the upper parts of the Proterozoic water column. In present-
day systems, sulfidic waters develop above sediments that
produce more sulfide than they consume by pyrite formation
and/or sulfide oxidation, and where additional production of
sulfide in the water column extends the depth interval of
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 523
0361-0128/98/000/000-00 $6.00 523
sulfidic waters (e.g., Yakushev and Neretin, 1997; Srensen
and Canfield, 2004). In todays environments, sulfidic waters
extend up to a chemocline where they are removed by oxida-
tion reactions and other Fe-O-S chemistry. In at least one
present-day system, sulfidic waters have developed in the
water column without a supply of sulfide from underlying
sediments. Crowe et al. (2008) documented a system in Lake
Matano, Indonesia, in which oxic waters overlie sulfidic wa-
ters that in turn overlie iron-rich waters. The controls on this
system appear to be the delivery of organic carbon and the
availability of water column sulfate and iron. Both of these
systems provide insights into the distribution of sulfidic, oxy-
genated, and ferrous waters in the Proterozoic oceans.
The surface waters of the Proterozoic oceans are thought to
have been oxygenated as a result of oxygenic photosynthesis.
Johnston et al. (2009) suggest that a sulfide-rich oxygen min-
imum zone would have played a critical role in limiting oxy-
gen production in the surface waters of the Proterozoic ocean
and a stabilizing feedback that ultimately determined the at-
mospheric oxidation state and the redox structure of the
oceans. Oxygen limitation in oceanic shelf settings where the
supply of organic electron donors for sulfate reduction was high
would lead to sulfide production. If the sulfide production
exceeded water column oxygen and iron supply, then euxinic
conditions would have developed. This may have occurred as
a result of sulfide released from sediments or from enhanced
sulfate reduction in the water column such as that observed
by Crowe et al. (2008). Support for this has been provided by
Poulton et al. (2004), who provide evidence from iron speci-
ation that points to the development of sulfidic waters in Pro-
terozoic shelf environments. Lower sulfate reduction rates
are observed today in abyssal sediments (~0.121 mol/cm
2
/
yr) than in shelf or slope sediments (~40270 and ~ 1574
mol/cm
2
/yr, respectively; Canfield et al. 2005). There is no
reason to suspect that a similar relative relationship between
sulfate reduction rates did not occur in the Proterozoic. This
may have limited the supply of sulfide from Proterozoic
abyssal sediments to abyssal waters, and it may have produced
a situation in which the competition between sulfide and
deep sources of iron may not have favored euxinia, allowing
for formation of iron oxides without appreciable iron sulfide
in a number of Proterozoic VMS deposits (Slack et al., 2007).
It is not clear whether intermediate waters were sulfidic, such
as those that develop in modern settings like that of Lake
Matano (Crowe et al., 2008), but this is a possibility that some
favor. We suggest that the Proterozoic oceans may have had a
524 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 524
TABLE 2. Observations Relevant to Proterozoic Ocean Chemistry
Type of evidence Age Implications for oceans Setting Reference
Evidence for oxygenated oceanic pools
Presence of hematite in the jasper and iron formation 1740 Ma, Oxygenated water Deep water 1, 2
Jerome deposit Since 1.85 Ga (>1,000 m)
Cerium abundances in Proterozoic jasper from deep-water 1740 Ma Low oxygen, Whole ocean 1
exhalites suggest limited formation of Mn and Fe-, Mn-oxides limited formation
of Fe,- Mn-oxide
Mo isotope evidence suggests that the global ocean chemistry 1.8 Ga 1.0 Ga Low oxygen/ Whole ocean 3, 4
did not preserve the same signal as seen in the oceans today Limited formation
that results from formation of Mn and Fe-, Mn-oxides of Fe,- Mn-oxide
Occurrence of non-sulfide zinc deposits such as the Franklin Late Low sulfide, possibly Below wave base 5
and the Sterling Hill deposits of New Jersey Mesoproterozoic low total sulfur.
Evidence for sulfidic oceanic pools
Lack of a barite cap on the Sullivan CD Pb-Zn deposit ~1430 Ma Low sulfate Below wave base 6
Evidence from Mo abundances and Mo-TOC suggest a ~1.8 Ga 551 Ma. Sulfidic waters or Whole ocean 7
significant sulfidic sink for Mo other Sulfidic sinks
Presence of pigment biomarkers from two classes of ~1.64 Ga Sulfidic waters Photic zone/ 8, 9, 10
phototrophic sulfide oxidizing bacteria (purple and green possible restricted
sulfur bacteria) in McArthur basin sediments marine basin
Change in iron speciation in the Gunflint Iron Formation and ~1.85 Ga Transition from ferrous Shelf below 11
the overlying Rove Fm. to sulfidic waters wave base
Models for formation of CD Pb-Zn deposits in the McArthur ~1.6 Ga Sulfidic bottom waters Below wave 6, 12
basin that include mixing of deep (oxidized brines) with base, possible
sulfidic deep ocean-derived shallow fluids restricted basin
Long-term, positive
34
S of Mesoproterozoic sulfate and sulfide 1.8 Ga 1.0 Ga Missing sulfidic sink Whole ocean Figures 1,
(average value) from sedimentary rocks and ore deposits implies sulfidic waters 810 and
suggests a missing negative
34
S pool since the ultimate and sediments in see 12, 13,
sources of sulfur to the surface sulfur cycle (input of juvenile deep water 14, 15
sulfur) have
34
S values close to zero
(1) Slack et al. (2007), (2) Slack and Cannon (2009), (3) Anbar and Knoll (2002), (4) Arnold et al. (2004), (5) Holland (2005), (6) Cooke et al. (2000), (7)
Scott et al. (2008), (8) Brocks et al. (2004), (9) Brocks et al. (2005), (10) Brocks and Schaeffer (2008), (11) Poulton et al. (2004), (12) Lyons et al. (2004), (13)
Logan et al. (1995), (14) Canfield (2004), (15) Johnston et al. (2007)
complex structure with respect to sulfidic, oxygenated, and
ferruginous waters. Whereas surface waters were oxygenated,
water over the shelves, the intermediate-water, and perhaps a
number of deep basins may have developed euxinic condi-
tions, and other deep-water settings may have been anoxic,
nonsulfidic, or even ferruginous.
Some features of the record preserved by ore deposits
and sedimentary geochemistry suggest an even more com-
plex picture. These include the following: (1) a suggestion
by Johnston et al. (2005a) that the oxidation pathways for
sulfide in the Proterozoic oceans underwent a change in the

33
S of sulfate that started at ~1.4 Ga, (2) the appearance
and disappearance of nonsulfide zinc deposits (Holland,
2005), (3) a drop in the frequency of CD Pb-Zn deposits in
this interval (Lyons et al., 2004), (4) the absence of signifi-
cant evidence for the presence of euxinic conditions be-
tween ~1.5 Ga and 750 Ma (Condie et al., 2001; Meyer and
Kump, 2008; Scott et al., 2008), and (5) the suggestion by
Canfield et al. (2008), on the basis of iron speciation, that
anoxic but ferruginous water replaced sulfidic waters in the
Neoproterozoic because of enhanced iron delivery or lim-
ited sulfur. These changes suggest something happened
with the sulfur chemistry of the oceans in the late Protero-
zoic. One explanation is that they reflect the limitation of
sulfur on a global scale brought about by a shrinking exo-
genic sulfur pool, a suggestion made by Canfield (2004) on
the grounds that sulfide deposited in abyssal sediments may
have been lost from the system to subduction. Another ex-
planation by Holland (2005) attributes these changes to
higher fluxes of sinking organic matter, exceeding the ven-
tilation (oxygenation) of the deep ocean that left Fe
2+
in ex-
cess. It is not clear if they can be linked to a single overriding
cause, but it would appear that further studies of ore de-
posits and sedimentary sulfur in combination with biogeo-
chemical models are needed to provide an answer to why
these changes have been suggested.
Synthesis
Evidence from ore deposits and the sedimentary rock
record yield several important observations on the connec-
tions between the sulfur cycle and the secular evolution of
ocean chemistry. The most significant of these are summa-
rized below and are used to develop an evolutionary model
for the principal features of the oceans (Fig. 12).
The presence of Superior-type banded iron formations
(Cloud, 1973) is evidence for significant solubility of ferrous
iron (Fe
2+
) in the oceans that limited oxygen and free sulfide.
On the basis of trace metal and sulfur isotope data, it has been
suggested that local, or spatially constrained oxygenated
(Wille et al., 2006; Anbar et al., 2007; Kaufman et al., 2007),
or sulfidic pools (Reinhard et al., 2009; Poulton et al., 2009)
developed in the Neoarchean ocean that reflected the ap-
pearance of oxygenic photosynthesizers. These suggestions
indicate that the oceans of early Earth evolved from a Pale-
oarchean and Mesoarchean state with ferruginous deep wa-
ters and relatively small amounts of cycling to a Neoarchean
state with ferruginous deep waters and surface waters with
production of oxygen and local competition between iron,
oxygen, and sulfur. Nevertheless, oceanic sulfate levels likely
remained below 200 M during the entire Archean (Habicht
et al., 2002), and there is no conclusive evidence for spikes in
sulfate concentration. This is consistent with the lack of sig-
nificant variability for
34
S and the preservation of MIF-S in
the sedimentary record.
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 525
0361-0128/98/000/000-00 $6.00 525
Sulfidic
Ferruginous
F
e
r
r
u
g
i
n
o
u
s
Oxic
Oxic
D
e
e
p

O
c
e
a
n

C
h
e
m
i
s
t
r
y
A
t
m
o
s
p
h
e
r
e
Surface
Ocean
Oxic Anoxic
low high
Ferruginous
O
c
e
a
n
i
c

s
u
l
f
a
t
e

c
o
n
c
e
n
t
r
a
t
i
o
n
Time (Ga)
3.8 2.8 2.42 1.8 1.4 0.8 0.55 0.3 Today
?
gypsum before halite halite before gypsum
bioturbation
OAEs
nonsulfidic
anoxic
FIG. 12. Inferred evolution of atmosphere, surface ocean, and deep ocean and oceanic sulfate concentration. Note that
the representation does not imply depth or geographic distribution of pools in the deep ocean.
Whereas some earlier studies of ore deposits have argued
for high levels of sulfate in the Paleoarchean (e.g., Vearn-
combe et al., 1995), this may not be correct (Huston and
Logan, 2004). It is possible, however, that the Neoarchean
may have been characterized by a drop in seawater sulfate
concentrations because of the growth of sinks associated
with microbial sulfate reduction. This is consistent with a
rise in the
34
S of the oceanic sulfate pool and possibly ex-
plains the lack of barite in Neoarchean VMS deposits. It also
may have contributed to development of euxinia in some
basins where sulfate reduction rates were enhanced, and if
the concentrations of sulfate exceeded the local supply of
ferrous iron.
The rise in oxygen levels at ~2.42 Ga was accompanied by
a rise in seawater sulfate concentrations, the shutting down
of MIF-S processes, an increase in the variability of
34
S,
and a relationship between
34
S, MIF-S, and
13
C. This
GOE was followed by a protracted change in the chemistry
of iron in the oceans, indicated by
56
Fe and Mo data
(Rouxel et al., 2005; Johnson et al., 2008b; Scott et al.,
2008). Other observations include a change in the sequence
of precipitation of gypsum and halite in evaporites
(Grotzinger and Kasting, 1993; Pope and Grotzinger, 2003)
that suggest sulfate and calcium concentrations exceeded a
saturation threshold of ~23 mM
2
at about 1.8 Ga. Quartz
nodules with palmate upright shapes and chevron termina-
tions consistent with a selenite gypsum precursor have
been reported in 2.1 Ga evaporites of the Lucknow For-
mation, South Africa (Schrder et al., 2008, p. 110), indicat-
ing that gypsum precipitation before halite occurred at an
even earlier date. There are other reports of higher sulfate
levels at 2.4 to 1.8 Ga, including evidence for gypsum pre-
cipitation (e.g., Mehlezhik et al., 2005). This time also
marked the appearance of the most ancient CD Pb-Zn de-
posits, MVT deposits and sediment-hosted copper deposits,
which all have been related to rising sulfate levels in near-
surface environments.
A change in ocean chemistry occurred at ~1.9 to 1.8 Ga,
when sulfidic and/or anoxic nonsulfidic, nonferruginous
ocean pools were established. The balance between sulfidic
and nonsulfidic anoxic pools in the 1.8 Ga to 800 Ma interval
was a balanced system that responded by changes in the pro-
portion of sulfidic to nonsulfidic anoxic pools.
Stabilization of sulfidic pools was favored in shelf settings
where sulfate delivery rates and sulfate reduction rates were
high, and also in intermediate waters below the photic zone
where delivery of organic donors was high (Crowe et al.,
2008). Development of nonsulfidic low oxygen pools oc-
curred where ventilation of the deep water exceeded the oxy-
gen sinks from settling organic matter and also from sources
of sulfide and ferrous iron. Evidence from CD Pb-Zn de-
posits in continental rifts and sag basins and from VMS de-
posits, as well as from sedimentary iron speciation, supports a
transition to an ocean with sulfidic as well as nonsulfidic and
anoxic (Slack et al., 2007; Slack and Cannon, 2009) interme-
diate and deep water at this time.
The second half of the Neoproterozoic, and possibly the
entire Neoproterozoic, was dominated by ferruginous ocean
bottom water in shelf settings (Canfield et al., 2008). The per-
sistence of some sulfidic pools is indicated by Mo data (Scott
et al., 2008), and the existence of shallow-water euxinia is in-
dicated by a biomarker (Olcott et al., 2005). The reasons for
the re-establishment of iron-rich conditions are debated, but
include hypotheses related to the size of the exogenic sulfur
pool (Canfield, 2004; Canfield et al., 2008) or the change in
the supply of hydrothermal ferrous iron.
Starting in the latest Neoproterozoic (~600 Ma), a second
oxidation of the Earths surface is indicated by the coloniza-
tion of the oceans by animal life (e.g., Logan et al., 1995). It
has also been suggested this change corresponded to a rise of
sulfate above Proterozoic levels that is attributable to the
onset of bioturbation (Canfield and Farquhar, 2009). Evi-
dence for this rise is indirectly tied to a change in the range of
variation for
34
S. Direct evidence for episodes of high sulfate
concentrations is present in rocks as old as latest Neoprotero-
zoic (Lowenstein et al., 2005) and in other parts of the geo-
logic record in the Phanerozoic (Horita et al., 1991, 1996,
2002; Lyons et al., 2004; Lowenstein et al., 2005; Fig. 7).
However, data do not provide clear evidence that sulfate lev-
els exceeded terminal Proterozoic values for all of the
Phanerozoic. We suggest that sulfate concentrations were,
however, generally higher in the Phanerozoic, which allowed
for increased formation of evaporates that provided favorable
conditions for generation of MVT deposits. The main control
on the temporal distribution of MVT deposits is the tectonic
remobilization of sulfate-rich evaporative concentrated sea-
water, rather than a maximum in oceanic sulfate concentra-
tion (see also Leach et al., 2010). It is unclear, however, how
the absence of continental rift- and sag-related CD Pb-Zn de-
posits are related to changing ocean chemistry (Leach et al.,
2010). This may reflect slightly higher oxygen levels in the
water column that inhibited formation of sulfidic conditions
in the sediments and in the deeper parts of these basins.
Some CD Pb-Zn deposits from this time suggest develop-
ment of euxinic conditions in certain basins (Goodfellow and
Jonasson, 1984).
The picture that has emerged is not simple, but it is con-
sistent with observations from a number of sources. Oceans
with a single, well-mixed surface pool and a well-mixed
deep-water pool were not present throughout most of the
geologic history. Rather, variations in the geographic and
temporal distribution of oxic, ferruginous, and sulfidic pools
occurred. The view that emerges is more intricate than de-
scribed by the existing models, but we do not believe that
we have added unnecessary complexity. Rather, we argue
that the combination of geochemically distinct pools is a
general feature of the oceans state during significant peri-
ods in Earth history, which is required by the contrasting ob-
servations of sulfidic, iron-rich, and oxygenic signals extend-
ing from the Neoarchean until the present. The balance
between these pools has varied considerably. Prior studies
have sought to understand ocean chemistry and its evolution
in geologic time in terms of the underlying controls on and
responses of the oceans to changes in the supply and sinks
of elements such as iron, oxygen, and sulfur. Future re-
search will be needed to understand the mechanisms that
drove large scale changes in ocean chemistry and to under-
stand the ways that feedbacks between elements of the sul-
fur cycle and responses of sulfur pools may have been
recorded in sediments and sulfide ore deposits.
526 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 526
Acknowledgments
JF acknowledges support from the Guggenheim Founda-
tion, The NASA Exobiology Program, NORDCEE, and the
Danish National Science foundation. Funding from the NSF
(EAR0918382) supported Nanping Wu during the final
stages of this project. Thoughtful reviews by Wayne Goodfel-
low, Craig Johnson, Rich Goldfarb, and David Leach as well
as a separate set of comments by Dick Holland were also
greatly appreciated. Collectively, these reviews provided crit-
ical and constructive comments resulted in significant
changes to the manuscript and are gratefully acknowledged.
REFERENCES
Akande, S.O., and Zentilli, M., 1984, Geologic, fluid inclusion, and stable iso-
tope studies of the Gays River lead-zinc deposit, Nova Scotia, Canada:
ECONOMIC GEOLOGY, v. 79, p. 11871211.
Algeo, T.J., and Lyons, T.W., 2006, Mo-total organic carbon covariation in
modern anoxic marine environments: Implications for analysis of paleore-
dox and paleohydrographic conditions: Paleoceanography, v. 21, PA1016
DOI: 10.1029/2004PA001112.
Anbar, A.D., and Knoll, A.H., 2002, Proterozoic ocean chemistry and evolu-
tion: A bioinorganic bridge?: Science, v. 297, p. 11371142.
Anbar, A.D., Duan, Y., Lyons, T.W., Arnold, G.L., Kendall, B., Creaser, R.A.,
Kaufman, A.J., Gordon, G.W., Scott, C., Garvin, J., and Buick, R., 2007, A
whiff of oxygen before the Great Oxidation Event?: Science, v. 317, p.
19031906.
Anderson, B.R., Gemmell, J.B., and Berry, R.F., 2001, The geology of the
Nifty copper deposit, Throssell Group, Western Australia: Implications for
ore genesis: ECONOMIC GEOLOGY, v. 96, p. 15351565.
Anderson, I.K., Ashton, J.H., Boyce, A.J., Fallick, A.E., and Russell, M.J.,
1998, Ore depositional processes in the Navan Zn-Pb deposit, Ireland:
ECONOMIC GEOLOGY, v. 93, p. 535563.
Andrew, A.S., Heinrich, C.A., Wilkins, R.W.T., and Patterson, D.J., 1989, Sul-
fur isotope systematics of copper ore formation at Mount Isa, Australia:
ECONOMIC GEOLOGY, v. 84, p. 16141626.
Anger, G., Nielsen, H., Puchelt, H., and Ricke, W., 1966, Sulfur isotopes in
the Rammelsberg ore deposit (Germany): ECONOMIC GEOLOGY, v. 61, p.
511536.
Arnold, G.L., Anbar, A.D., Barling, J., and Lyons, T.W., 2004, Molybdenum
isotope evidence for widespread anoxia in mid-Proterozoic oceans: Sci-
ence, v. 304, p. 8790.
Arnold, M. Bernard A.J., and Soler E., 1977, Premier Apport de la Geochi-
mie des Isotopes du Soufre la Comprehension de la Genese des Minera-
lisations Pyriteuses de la Province de Huelva (Espagne): Mineralium De-
posita, v. 12, p. 197218
Bao, H.M., Rumble, D., and Lowe, D.R., 2007, The five stable isotope com-
positions of Fig Tree barites: Implications on sulfur cycle in ca. 3.2 Ga
oceans: Geochimica et Cosmochimica Acta, v. 71, p. 48684879.
Bechtel, A., Shieh, Y-N., Pervaz, M., and Puttman, W., 1996, Biodegradation
of hydrocarbons and biogeochemical sulfur cycling in the salt dome envi-
ronment; inferences from sulfur isotope and organic geochemical investi-
gations of the Bahloul Formation at the Bou Grine Zn/Pb ore deposit:
Geochimica et Cosmochimica Acta, v. 60, p. 28332855.
Bekker, A., Holland, H.D., Wang, P.L., Rumble, D., Stein, H.J., Hannah,
J.L., Coetzee, L.L., and Beukes, N.J., 2004, Dating the rise of atmospheric
oxygen: Nature, v. 427, p. 117120.
Beukes, N.J., Klein, C., Kaufman, A.J., and Hayes, J.M., 1990, Carbonate
petrography, kerogen distribution, and carbon and oxygen isotope varia-
tions in an Early Proterozoic transition from limestone to iron-formation
deposition, Transvaal Supergroup, South-Africa: ECONOMIC GEOLOGY, v.
85, p. 663690.
Bhattacharya, S.K., Savarino, J., and Thiemens, M.H., 2000, A new class of
oxygen isotopic fractionation in photodissociation of carbon dioxide: Poten-
tial implications for atmospheres of Mars and Earth: Geophysical Research
Letters, v. 27, p. 14591462.
Blakeman, R.J., Ashton, J.H., Boyce, A.J., Fallick, A.E., and Russel, M.J.,
2002, Timing of interplay between hydrothermal and surface fluids in the
Navan Zn + Pb orebody, Ireland: Evidence from metal distribution
trends, mineral textures, and
34
S analyses: ECONOMIC GEOLOGY, v. 97, p.
7391.
Broadbent, G.C., Myers, R.E., and Wright, J.V., 1998, Geology and origin of
shale-hosted Zn-Pb-Ag mineralization at the Century deposit, northwest
Queensland, Australia: ECONOMIC GEOLOGY, v. 93, p. 12641294.
Brocks, J.J., and Schaeffer, P., 2008, Okenane, a biomarker for purple sulfur
bacteria (Chromatiaceae), and other new carotenoid derivatives from the
1640 Ma Barney Creek Formation: Geochimica et Cosmochimica Acta, v.
72, p. 13961414.
Brocks, J.J., Love, G.D., Summons, R.E., and Logan, G.A., 2004, Purple sul-
fur bacteria in an intensely stratified Paleoproterozoic sea: Geochimica et
Cosmochimica Acta, v. 68, p. A796A796.
Brocks, J. Love, G.D., Summons, R.E., Knoll, A.H., Logan, G.A., and Bow-
den, S.A., 2005, Biomarker evidence for green and purple sulphur bacteria
in a stratified Palaeoproterozoic sea: Nature, v. 437, p. 866870.
Brown, A.C., 1997, World-class sediment-hosted stratiform copper deposits:
Characteristics, genetic concepts and metallotects: Australian Journal of
Earth Sciences, v. 44, p. 317328.
2005, Refinements for footwall red-bed diagenesis in the sediment-
hosted stratiform copper deposits model: ECONOMIC GEOLOGY, v. 100, p.
765771.
2009, A process-based approach to estimating the copper derived from
red beds in the sediment-hosted stratiform copper deposit model: ECO-
NOMIC GEOLOGY, v. 104, p. 857868.
Buchachenko, A.L., 1995, Mie versus Ciecomparative-analysis of magnetic
and classical isotope effects: Chemical Reviews, v. 95, p. 25072528.
Buick, R., and Dunlop, J.S.R., 1990, Evaporitic sediments of Early Archean
age from the Warrawoona Group, North-Pole, Western-Australia: Sedi-
mentology, v. 37, p. 247277.
Buick, R., Brauhart, C.W., Morant, P., Thornett, J.R., Maniw, J.G., Archibald,
N.J., Doepel, M.G., Fletcher, I.R., Pickard, A.L., Smith, J.B., Barley, M.E.,
McNaughton, N.J., and Groves, D.I., 2002, Geochronology and strati-
graphic relationships of the Sulphur Springs Group and Strelley Granite: A
temporally distinct igneous province in the Archaean Pilbara craton, Aus-
tralia: Precambrian Research, v. 114, p. 87120.
Buschendorf, Fr., Nielsen, H., Puchelt, H., and Ricke, W., 1963, Schwefe-
lisotopen-untersuchungen am pyrit-sphalerit-baryt-lager Meggen/Lenne
(Deutschland) und an verschiedenen Devon-evaporiten: Geochimica et
Cosmochimica Acta, v. 27, p. 501523.
Cagatay, M.N., and Eastoe, C. J., 1995, A sulfur isotope study of volcanogenic
massive sulfide deposits of the eastern Black-Sea province, Turkey: Miner-
alium Deposita, v. 30, p. 5566.
Cailteux, J.L.H., Kampunzu, A.B., Lerouge, C., Kaputo, A.K., and Milesi,
J.P., 2005, Genesis of sediment-hosted stratiform copper-cobalt deposits,
central African Copperbelt: Journal of African Earth Sciences, v. 42, p.
134158.
Cameron, E.M., 1982, Sulfate and sulfate reduction in early Precambrian
oceans: Nature, v. 296, p. 145148.
1983, Evidence from Early Proterozoic anhydrite for sulfur isotopic par-
titioning in Precambrian oceans: Nature, v. 304, p. 5456.
Cameron, E.M., and Hattori, K., 1987, Archean sulfur cycleevidence from
sulfate minerals and isotopically fractionated sulfides in Superior province,
Canada: Chemical Geology, v. 65, p. 341358.
Campbell, F.A., Ethier, V.G., Krouse, H.R., and Both, R.A., 1978, Isotopic
composition of sulfur in Sullivan orebody British Columbia: ECONOMIC
GEOLOGY, v. 73, p. 246268.
Campbell, F.A., Ethier, V.G., and Krouse, H.R., 1980, The massive sulfide
zone: Sullivan orebody: ECONOMIC GEOLOGY, v. 75, p. 916926.
Canet, C., Alfonso, P., Melgarejo, J.C., and Belyatsky, B.V., 2004, Geochem-
ical evidences of sedimentary-exhalative origin of the shale-hosted PGE-
Ag-Au-Zn-Cu occurrences of the Prades Mountains (Catalonia, Spain):
Trace-element abundances and Sm-Nd isotopes: Journal of Geochemical
Exploration, v. 82, p. 1733.
Canfield, D.E., 1998, A new model for Proterozoic ocean chemistry: Nature,
v. 396, p. 450453.
2001, Biogeochemistry of sulfur isotopes: Reviews in Mineralogy and
Geochemistry, v. 43, p. 607636.
2004, The evolution of the Earth surface sulfur reservoir: American
Journal of Science, v. 304, p. 839861.
Canfield, D.E., and Farquhar, J., 2009, Animal evolution, bioturbation, and
the sulfate concentration of the oceans: Proceedings of the National Acad-
emy of Science, v. 106, p. 81238127.
Canfield, D.E., and Teske, A., 1996, Late Proterozoic rise in atmospheric
oxygen concentration inferred from phylogenetic and sulphur-isotope stud-
ies: Nature, v. 382, p. 127132.
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 527
0361-0128/98/000/000-00 $6.00 527
Canfield, D.E., Habicht, K.S., and Thamdrup, B., 2000, The Archean sulfur
cycle and the early history of atmospheric oxygen: Science, v. 288, p.
658661.
Canfield, D.E., Thamdrup, B., Kristensen, E., 2005, Aquatic geomicrobiol-
ogy: Advances in Marine Biology, v. 48, San Diego, California, Elsevier, 640
p.
Canfield, D.E., Poulton, S.W., Knoll, A.H., Narbonne, G.M., Ross, G., Gold-
berg, T., and Strauss, H., 2008, Ferruginous conditions dominated later
Neoproterozoic deep-water chemistry: Science, v. 321, p. 949952.
Carr, G.R., and Smith, J.W., 1977, A comparative isotopic study of the Lady
Loretta zinc-lead-silver deposit: Mineralium Deposita, v. 12, p. 105110.
Cates, N.L., and Mojzsis, S.J., 2006, Chemical and isotopic evidence for
widespread Eoarchean metasedimentary enclaves in southern West Green-
land: Geochimica et Cosmochimica Acta, v. 70, p. 42294257.
Claypool, G.E., Holser, W.T., Kaplan, I.R., Sakai, H., and Zak, I., 1980, The
age curves of sulfur and oxygen isotopes in marine sulfate and their mutual
interpretation: Chemical Geology, v. 28, p. 199260.
Cloud, P., 1973, Paleoecological significance of banded iron-formation: ECO-
NOMIC GEOLOGY, v. 68, p. 11351143.
Condie, K.C., Marais, D.J.D., and Abbott, D., 2001, Precambrian super-
plumes and supercontinents: A record in black shales, carbon isotopes, and
paleoclimates?: Precambrian Research, v. 106, p. 239260.
Cook, N.J., and Hoefs, J., 1997, Sulphur isotope characteristics of metamor-
phosed Cu-(Zn) volcanogenic massive sulphide deposits in the Norwegian
Caledonides: Chemical Geology, v. 135, 307324.
Cooke, D.R., Bull, S.W., Large, R.R., and McGoldrick, P. ., 2000, The im-
portance of oxidized brines for the formation of Australian proterozoic
stratiform sediment-hosted Pb-Zn (sedex) deposits: ECONOMIC GEOLOGY,
v. 95, p. 117.
Craig, H., 1957, Isotopic standards for carbon and oxygen and correction fac-
tors for mass-spectrometric analysis of carbon dioxide: Geochimica et Cos-
mochimica Acta, v. 12, p. 133149.
Crocetti, C.A., and Holland, H.D., 1989, Sulfur-lead isotope systematics and
the composition of fluid inclusions in galena from the Viburnum Trend,
Missouri: ECONOMIC GEOLOGY, v. 84, p. 21962216.
Crowe, S.A., Jones, C., Katsev, S., Magen, C., ONeill, A.H., Sturm, A., Can-
field, D.E., Haffner, G.D., Mucci, A., Sundby, B., and Fowle, D.A., 2008,
Photoferrotrophs thrive in an Archean Ocean analogue: Proceedings of the
National Academy of Sciences of the United States of America, v. 105, p.
15,93815,943.
Davidson, G.J., and Dixon, G.H., 1992, Two sulphur isotope provinces de-
duced from ores in the Mount Isa eastern succession, Australia: Mineral-
ium Deposita, v. 27, p. 3041.
DeChow, E, 1966: Geology, sulfur isotopes and the origin of the Heath
Steele ore deposits, Newcastle, N. B., Canada:, ECONOMIC GEOLOGY; v. 55,
p. 539556.
Decre, S., Marignac, C., De Putter, T., Deloule, E., Liegeois, J.P., and De-
maiffe, D., 2008, Pb-Zn mineralization in a Miocene regional extensional
context: The case of the Sidi Driss and the Douahria ore deposits (Nefza
mining district, northern Tunisia): Ore Geology Reviews, v. 34, p. 285303.
Deloule, E., Allegre, C., and Doe, B., 1986, Lead and sulfur isotope micros-
tratigraphy in galena crystals from Mississippi Valley-style deposits:
ECONOMIC GEOLOGY, v. 81, p. 13071321.
Ding, T.-P., and Jiang, S.-Y., 2000, Stable isotope study of the Langshan poly-
metallic mineral district, Inner Mongolia, China: Resource Geology, v. 50,
p. 2538.
Dixon, G.H., and Davidson, G.J., 1996, Stable isotope evidence for thermo-
chemical sulfate reduction in the Dugald River (Australia) strata-bound
shale-hosted zinc-lead deposit: Chemical Geology, v. 129, p. 227246.
Domagal-Goldman, S.D., Kasting, J.F., Johnston, D.T., and Farquhar, J.,
2008, Organic haze, glaciations and multiple sulfur isotopes in the Mid-
Archean Era: Earth and Planetary Science Letters, v. 269, p. 2940.
Eastoe, C.J., and Nelson S.E., 1988, A Permian kuroko-type hydrothermal
system, Afterthought-Ingot area, Shasta County, California; lateral and ver-
tical sections, and geochemical evolution: ECONOMIC GEOLOGY; v. 83; p.
588605.
Eldridge, C.S., Williams, N., and Walshe, J.L., 1993, Sulfur isotope variabil-
ity in sediment-hosted massive sulfide deposits as determined using the ion
microprobe SHRIMP: II. A study of the H.Y.C. Deposit at McArthur River,
Northern Territory, Australia: ECONOMIC GEOLOGY, v. 88, p. 126.
Farquhar, J., and Johnston, D.T., 2008, The oxygen cycle of the terrestrial
planets: Insights into the processing and history of oxygen in surface envi-
ronments, in MacPherson, G.J., Mittlefehldt, D.W., Jones, J.H., and Simon,
S.B., eds., Oxygen in the Solar System: Reviews in Mineralogy and Geo-
chemistry, v. 68, p. 463492.
Farquhar, J., and Wing, B.A., 2003, Multiple sulfur isotopes and the evolution
of the atmosphere: Earth and Planetary Science Letters, v. 213, p. 113.
2005, The terrestrial record of stable sulphur isotopes: A review of the
implications for evolution of Earths sulphur cycle, in McDonald, I., Boyce,
A.J., Butler, I.B.,Herrington, R.J. and Polya, D.A., eds., Mineral Deposits
and Earth Evolution. Geological Society, London, Special Publications, v.
248, p. 167177.
Farquhar, J., Bao, H.M., and Thiemens, M., 2000, Atmospheric influence of
Earths earliest sulfur cycle: Science, v. 289, p. 756758.
Farquhar, J., Savarino, J., Airieau, S., and Thiemens, M.H., 2001, Observa-
tion of wavelength-sensitive mass-independent sulfur isotope effects dur-
ing SO2 photolysis: Implications for the early atmosphere: Journal of Geo-
physical Research-Planets, v. 106, p. 32,82932,839.
Farquhar, J., Wing, B.A., McKeegan, K.D., Harris, J.W., Cartigny, P., and
Thiemens, M.H., 2002, Mass-independent sulfur of inclusions in diamond
and sulfur recycling on early earth: Science, v. 298, p. 23692372.
Farquhar, J., Peters, M., Johnston, D.T., Strauss, H., Masterson, A., Wiechert,
U., and Kaufman, A J., 2007a, Isotopic evidence for Mesoarchaean anoxia
and changing atmospheric sulphur chemistry: Nature, v. 449, p. 706U5.
(p. 706709 and digital supplement).
Farquhar, J., Johnston, D.T., and Wing, B.A., 2007b, Implications of conser-
vation of mass effects on mass-dependent isotope fractionations: Influence
of network structure on sulfur isotope phase space of dissimilatory sulfate
reduction: Geochimica et Cosmochimica Acta, v. 71, p. 58625875.
Foriel, J., Philippot, P., Rey, P., Somogyi, A., Banks, D., and Menez, B., 2004,
Biological control of Cl/Br and low sulfate concentration in a 3.5-Gyr-old
seawater from North Pole, Western Australia: Earth and Planetary Science
Letters, v. 228, p. 451463.
Frietsch, R., Billstrom, K., Perdahl, Ja., 1995, Sulfur Isotopes in lower Pro-
terozoic iron and sulfide ores in northern Sweden: Mineralium Deposita,
v.30, p. 275284.
Gaboury, D., Daigneault, R., and Beaudoin, G., 2000, Volcanogenic-related
origin of sulfide rich quartz veins: Evidence from O and S isotopes at the
Giant dormant gold mine, Abitibi belt, Canada: Mineralium Deposita, v.
35, p. 2136.
Gao, Y.Q., and Marcus, R.A., 2001, Strange and unconventional isotope ef-
fects in ozone formation: Science, v. 293, p. 259263.
Gardner, H.D., and Hutcheon, I., 1985, Geochemistry, mineralogy, and ge-
ology of the Jason Pb-Zn deposits, Macmillan Pass, Yukon, Canada:
ECONOMIC GEOLOGY, v. 80, p. 12571276.
Gellatly, A.M., and Lyons, T. W., 2005, Trace sulfate in mid-Proterozoic car-
bonates and the sulfur isotope record of biospheric evolution: Geochimica
et Cosmochimica Acta, v. 69, p. 38133829.
Gilg, H.A., Boni, M., Balassone, G., Allen, C.R., Banks, D., and Moore, F.,
2006, Marble-hosted sulfide ores in the Angouran Zn-(Pb-Ag) deposit, NW
Iran: interaction of sedimentary brines with a metamorphic core complex:
Mineralium Deposita, v. 41, p. 116.
Golding, S.D., and Young, E., 2005, Multiple sulfur isotope evidence for dual
sulfur sources in the 3.24 Ga Sulphur Springs VHMS deposit: Geochimica
et Cosmochimica Acta, v. 69, p. A449.
Golding, S.D., Young, E., Duck, L.J., Baublys, K.A., and Glikson, M., 2006,
Multiple sulfur isotope constraints on microbial processes in Archaean
seafloor environments: Geochimica et Cosmochimica Acta, v. 70, p. A208.
Goodfellow, W.D., 1987, Anoxic stratified oceans as a source of sulfur in sed-
iment-hosted stratiform Zn-Pb deposits (Selwyn basin, Yukon, Canada):
Chemical Geology, v. 65, p. 359382.
Goodfellow, W.D., and Jonasson, I.R., 1984, Ocean stagnation and ventilation
defined by Delta-S-34 secular trends in pyrite and barite, Selwyn basin,
Yukon: Geology, v. 12, p. 583586.
Goodfellow, W.D., and Peter, J.M., 1996, Sulphur isotope composition of the
Brunswick No 12 massive sulphide deposit, Bathurst Mining Camp, New
Brunswick: Implications for ambient environment, sulphur source, and ore
genesis: Canadian Journal of Earth Sciences, v. 33, p. 231251.
Goodfellow, W.D., Lydon, J.W., and Turner, R.J.W., 1993: Geology and gen-
esis of stratiform sediment-hosted (SEDEX) zinc-lead-silver sulfide de-
posits, in Kirkham, R.V., Sinclair, W.D., Thorpe, R.I., and Duke, J.M., eds.:
Mineral deposit modeling: Geological Association of Canada, Special Paper
40, p. 201251.
Green, G.R., Solomon, M., and Walshe J.L., 1981, The formation of the vol-
canic-hosted massive sulfide ore deposit at Rosebery, Tasmania: ECONOMIC
GEOLOGY, v. 76, p. 304338.
528 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 528
Grotzinger, J.P., and Kasting, J.F., 1993, New constraints on Precambrian
ocean composition: Journal of Geology, v. 101, p. 235243.
Guha, J., 1971, Sulfur isotope study of the pyrite deposit of Amjhore, Shah-
bad district, Bihar, India: ECONOMIC GEOLOGY, v. 66, p. 326330.
Guo, Q., Strauss, H., Schrder, S., Gutzmer, J., Wing, B.A., Baker, M.A.,
Kaufman, A.J., Kim, S.T, and Farquhar, J., 2009, Reconstructing Earths
surface oxidation across the Archean-Proterozoic transition: Geology, v.37,
p. 399402.
Habicht, K.S., Gade, M., Thamdrup, B., Berg, P., and Canfield, D.E., 2002,
Calibration of sulfate levels in the Archean Ocean: Science, v. 298, p.
23722374.
Halverson, G.P., and Hurtgen, M.T., 2007, Ediacaran growth of the marine
sulfate reservoir: Earth and Planetary Science Letters, v. 263, p. 3244.
Haqq-Misra, J.D., Domagal-Goldman, S.D., Kasting, P.J., and Kasting, J.F.,
2008, A revised, hazy methane greenhouse for the Archean Earth: Astrobi-
ology, v. 8, p. 11271137.
Harrison, A.G., and Thode, H.G., 1958, Mechanism of the bacterial reduc-
tion of sulphate from isotope fractionation studies: Transactions of the
Faraday Society, v. 54, p. 8492.
Hathorn, B.C., and Marcus, R.A., 2000, An intramolecular theory of the
mass-independent isotope effect for ozone. II. Numerical implementation
at low pressures using a loose transition state: Journal of Chemical Physics,
v. 113, p. 94979509.
Heiligmann, M., Williams-Jones, A.E., and Clark, J.R., 2008, The role of sul-
fate-sulfide-oxide-silicate equilibria in the metamorphism of hydrothermal
alteration at the Hemlo gold deposit, Ontario: ECONOMIC GEOLOGY, v. 103,
p. 335351.
Heinrichs, T.K., and Reimer, T.O., 1977, Sedimentary barite deposit from
Archean Fig Tree Group of Barberton Mountain land (South Africa): ECO-
NOMIC GEOLOGY, v. 72, p. 14261441.
Hitzman, M., Kirkham, R., Broughton, D., Thorson, J., and Selley, D., 2005,
The sediment-hosted stratiform copper ore system: ECONOMIC GEOLOGY
100
TH
ANNIVERSARY VOLUME, p. 609642.
Hitzman, M.W., Selley, D., and Bull, S., 2010, Formation of sedimentary
rock-hosted stratiform copper deposits through Earth history: ECONOMIC
GEOLOGY, v. 105, p. 627639.
Hoering, T.C., 1989, The isotopic composition of bedded barites from the
Archean of southern India: Journal of the Geological Society of India, v. 34,
p. 461466.
Holland, H.D., 1984, The chemical evolution of the atmosphere and oceans:
Princeton Series in Geochemistry, 598 p.
2005, Sedimentary mineral deposits and the evolution of earths near-
surface environments: ECONOMIC GEOLOGY, v. 100, p. 14891509.
2006, The oxygenation of the atmosphere and oceans: Philosophical
Transactions of The Royal Society B-Biological Sciences, v. 361, p.
903915.
Horita, J., Friedman, T.J., Lazar, B., and Holland, H.D., 1991, The composi-
tion of Permian seawater: Geochimica et Cosmochimica Acta, v. 55, p.
417432.
Horita, J., Weinberg, A., Das, N., and Holland, H.D., 1996, Brine inclusions
in halite and the origin of the Middle Devonian Prairie evaporites of west-
ern Canada: Journal of Sedimentary Research, v. 66, p. 956964.
Horita, J., Zimmermann, H., and Holland, H.D., 2002, Chemical evolution
of seawater during the Phanerozoic: Implications from the record of ma-
rine evaporites: Geochimica et Cosmochimica Acta, v. 66, p. 37333756.
Hu, G.X., Rumble, D., and Wang, P.L., 2003, An ultraviolet laser microprobe
for the in situ analysis of multisulfur isotopes and its use in measuring
Archean sulfur isotope mass-independent anomalies: Geochimica et Cos-
mochimica Acta, v. 67, p. 31013118.
Huizenga, J.M., Gutzmer, J., Greyling, L.N., and Schaefer, M., 2006, Car-
bonic fluid inclusions in Paleoproterozoic carbonate-hosted Zn-Pb deposits
in Griqualand West, South Africa: South African Journal of Geology, v. 109,
p. 5562.
Hurtgen, M.T., Arthur, M.A., and Halverson, G.P., 2005, Neoproterozoic sul-
fur isotopes, the evolution of microbial sulfur species, and the burial effi-
ciency of sulfide as sedimentary pyrite: Geology, v. 33, p. 4144.
Hurtgen, M.T., Arthur, M.A., Suits, N.S., and Kaufman, A. J., 2002, The sul-
fur isotopic composition of Neoproterozoic seawater sulfate: Implications
for a snowball Earth?: Earth and Planetary Science Letters, v. 203, p.
413429.
Huston, D.L., and Logan, G.A., 2004, Barite, BIFs and bugs: Evidence for
the evolution of the Earths early hydrosphere: Earth and Planetary Science
Letters, v. 220, p. 4155.
Huston, D.L., Sun, S.S., Blewett, R., Hickman, A.H., Van Kranendonk, M.,
Phillips, D., Baker, D., and Brauhart, C., 2002, The timing of mineraliza-
tion in the archean North Pilbara terrain, Western Australia: ECONOMIC
GEOLOGY, v. 97, p. 733755.
Huston, D.L., Stevens, B., Southgate, P.N., Muhling, P., and Wyborn, L.,
2006, Australian Zn-Pb-Ag ore-forming systems: A review and analysis:
ECONOMIC GEOLOGY, v. 101, p. 11171157.
Huston, D.L., Pehrsson, S., Eglington, B.M., and Zaw, K., 2010, The geology
and metallogeny of volcanic-hosted massive sulfide deposits: Variations
through geologic time and with tectonic setting: ECONOMIC GEOLOGY, v.
105, p. 571591.
Jamieson, J.W., Wing, B.A., Hannington, M.D., and Farquhar, J., 2006, Eval-
uating isotopic equilibrium among sulfide mineral pairs in Archean ore de-
posits: Case study from the Kidd Creek VMS deposit, Ontario, Canada:
ECONOMIC GEOLOGY, v. 101, p. 10551061.
Johnson, C.A., and Emsbo, P., 2005, Possible impact of hydrothermal brine
discharge events on sulfur isotopes in marine basins and the global oceans:
Earth Systems Processes, v. 2, no. 1, p. 55.
Johnson, C.A., Emsbo, P., Poole, F.G., and Rye, R.O., 2009, Sulfur- and oxy-
gen-isotopes in sediment-hosted stratiform barite deposits: Geochimica et
Cosmochimica Acta, v. 73, p. 133147.
Johnson, C.M., Beard, B.L., Klein, C., Beukes, N.J., and Roden, E.E., 2008b,
Iron isotopes constrain biologic and abiologic processes in banded iron for-
mation genesis: Geochimica et Cosmochimica Acta, v. 72, p. 151169.
Johnston, D.T., Wing, B.A., Farquhar, J., Kaufman, A.J., Strauss, H., Lyons,
T.W., Kah, L.C., and Canfield, D.E., 2005a, Active microbial sulfur dispro-
portionation in the Mesoproterozoic: Science, v. 310, p. 14771479.
Johnston, D.T., Farquhar, J., Wing, B.A., Kaufman, A., Canfield, D.E., and
Habicht, K.S., 2005b, Multiple sulfur isotope fractionations in biological
systems: A case study with sulfate reducers and sulfur disproportionators:
American Journal of Science, v. 305, p. 645660.
Johnston, D.T., Poulton, S.W., Fralick, P.W., Wing, B.A., Canfield, D.E., and
Farquhar, J., 2006, Evolution of the oceanic sulfur cycle at the end of the
Paleoproterozoic: Geochimica et Cosmochimica Acta, v. 70, p. 57235739.
Johnston, D.T., Farquhar, J., and Canfield, D.E., 2007, Sulfur isotope in-
sights into microbial sulfate reduction: When microbes meet models:
Geochimica et Cosmochimica Acta, v. 71, p. 39293947.
Johnston, D.T., Farquhar, J., Summons, R.E., Shen, Y., Kaufman, A.J., Mas-
terson, A.L., and Canfield, D.E., 2008, Sulfur isotope biogeochemistry of
the Proterozoic McArthur Basin: Geochimica et Cosmochimica Acta, v. 72,
p. 42784290.
Johnston, D.T., Wolfe-Simon, F., Pearson, A., and Knoll, A. H., 2009, Anoxy-
genic photosynthesis modulated Proterozoic oxygen and sustained Earths
middle age: Proceedings of the National Academy of Sciences of the
United States of America, v. 106, p. 1692516929.
Jones, H.D., Kesler, S.E., Furman, F.C., and Kyle, J.R., 1996, Sulfur isotope
geochemistry of southern Appalachian Mississippi Valley-type deposits:
ECONOMIC GEOLOGY, v. 91, p. 355367.
Kah, L.C., Lyons, T.W., and Frank, T.D., 2004, Low marine sulphate and pro-
tracted oxygenation of the Proterozoic biosphere: Nature, v. 431, p. 834838.
Kajiwara Y., 1971, Sulfur isotope study of the Kuroko-ores of the Shakanai
no. 1 deposits, Akita Prefecture, Japan: Geochemical Journal, v. 4, p.
157181.
Kakegawa, T., Kawai, H., and Ohmoto, H., 1998, Origins of pyrites in the
similar to 2.5 Ga Mt. McRae Shale, the Hamersley district, Western Aus-
tralia: Geochimica et Cosmochimica Acta, v. 62, p. 32053220.
Kamber, B.S., and Whitehouse, M.J., 2007, Micro-scale sulphur isotope evi-
dence for sulphur cycling in the late Archean shallow ocean: Geobiology, v.
5, p. 517.
Kampschulte, A., and Strauss, H., 2004, The sulfur isotopic evolution of
Phanerozoic seawater based on the analysis of structurally substituted sul-
fate in carbonates: Chemical Geology, v. 204, p. 255286.
Karhu, J.A., and Holland, H.D., 1996, Carbon isotopes and the rise of at-
mospheric oxygen: Geology, v. 24, p. 867870.
Kase, K., Yamamoto, M., Nakamura, T., and Mitsuno, C., 1990, Ore miner-
alogy and sulfur isotope study of the massive sulfide deposit of Filon-Norte,
Tharsis mine, Spain: Mineralium Deposita, v. 25, p. 289296.
Kasting, J.F., 2001, Earth historythe rise of atmospheric oxygen: Science,
v. 293, p. 819820.
Kaufman, A.J., Johnston, D.T., Farquhar, J., Masterson, A.L., Lyons, T.W.,
Bates, S., Anbar, A.D., Arnold, G.L., Garvin, J., and Buick, R., 2007, Late
Archean biospheric oxygenation and atmospheric evolution: Science, v.
317, p. 19001903.
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 529
0361-0128/98/000/000-00 $6.00 529
Kelley, K.D., Leach, D.L., Johnson, C.A., Clark, J.L., Fayek, M., Slack, J.F.,
Anderson, V.M., Ayuso, R.A., Ridley, W.I., 2004, Textural, compositional,
and sulfur isotope variations of sulfide minerals in the Red Dog Zn-Pb-Ag
deposits, Brooks Range, Alaska: Implications for ore formation: ECONOMIC
GEOLOGY, v. 99, p. 15091532.
Kesler, S.E., and Reich, M.L., 2006, Precambrian Mississippi Valley-type de-
posits; relation to changes in composition of the hydrosphere and atmos-
phere, in Reich, M.H., ed., Geological Society of America, Memoir 198, p.
185204.
Kesler, S.E., Appold, M.S., Martini, A.M., Walter, L.M., Huston, T.J., and
Kyle, J.R., 1995, Na-Cl-Br Systematics of mineralizing brines in Mississippi
Valley-type deposits: Geology, v. 23, p. 641644.
Kesler, S.E., Friedman, G.M., and Krstic, D., 1997, Mississippi Valley-type
mineralization in the Silurian paleoaquifer, central Appalachians: Chemical
Geology, v. 138, p. 127134.
Kim S-T., Farquhar J., Gutzmer J., and Kesler S., 2009, Sulfur isotope sys-
tematics of the Paleoproterozoic Bushy Park and Pering MVT deposits,
Geochimica et Cosmochimica Acta, v. 73, A654, supplement.
Knoll, A.H., and Beukes, N. J., 2009, Introduction: Initial investigations of a
Neoarchean shelf margin-basin transition (Transvaal Supergroup, South
Africa): Precambrian Research, v. 169, p. 114.
Kurtz, A.C., Kump, L.R., Arthur, M.A., Zachos, J.C., and Paytan, A., 2003,
Early Cenozoic decoupling of the global carbon and sulfur cycles: Paleo-
ceanography, v. 18, DOI: 10.1029/2003PA000908.
Large, R.R., Bull, S.W., Cooke, D.R., and McGoldrick, P.J., 1998, A genetic
model for the HYC deposit, Australia: Based on regional sedimentology,
geochemistry, and sulfide-sediment relationships: ECONOMIC GEOLOGY, v.
93, p. 13451368.
Lasaga, A.C., Otake, T., Watanabe, Y., and Ohmoto, H., 2008, Anomalous
fractionation of sulfur isotopes during heterogeneous reactions: Earth and
Planetary Science Letters, v. 268, p. 225238.
Leach, D.L., Dwight, B., Lewchuk, M.T., Symons, D T.A., de Marsily, G.,
and Brannon, J., 2001, Mississippi Valley-type lead-zinc deposits through
geological time: implications from recent age-dating research: Mineralium
Deposita, v. 36, p. 711740.
Leach, D.L. Sangster, D.F., Kelley, K.D. Large, R.R, Garven, G. Allen, C.R.,
Gutzmer, J., and Walters, W., 2005, Sediment-hosted lead-sink deposits: A
global perspective: ECONOMIC GEOLOGY, 100
TH
ANNIVERSARY VOLUME, p.
561607.
Leach, D., Macquar, J.C., Lagneau, V., Leventhal, J., Emsbo, P., and Premo,
W., 2006, Precipitation of lead-zinc ores in the Mississippi Valley-type deposit
at Treves, Cevennes region of southern France: Geofluids, v. 6, p. 2444.
Leach, D.L., Bradley, D.C., Huston, D., Pisarevsky, S.A., Taylor, R.D., and
Gardoll, S.J., 2010, Sediment-hosted lead-zinc deposits in Earth history:
ECONOMIC GEOLOGY, v. 105, p. 593625.
Logan, G.A., Hayes, J.M., Hieshima, G.B., and Summons, R.E., 1995, Terminal
Proterozoic reorganization of biogeochemical cycles: Nature, v. 376, p. 5356.
Lott, D.A., 1999, Sedimentary exhalative nickel-molybdenum ores in South
China: ECONOMIC GEOLOGY, v. 94, p. 10511066.
Lowenstein, T.K., Timofeeff, M.N., Kovalevych, V.M., and Horita, J., 2005,
The major-ion composition of Permian seawater: Geochimica et Cos-
mochimica Acta, v. 69, p. 17011719.
Luepke, J.J., and Lyons, T.W., 2001, Pre-Rodinian (Mesoproterozoic) super-
continental rifting along the western margin of Laurentia: Geochemical ev-
idence from the Belt-Purcell Supergroup: Precambrian Research, v. 111, p.
7990.
Lusk, J., 1972, Examination of volcanic-exhalative and biogenic origins for
sulfur in the stratiform massive sulfide deposits of New Brunswick: ECO-
NOMIC GEOLOGY, v. 67, p. 169183.
Lyons, T.W., Luepke, J.J., Schreiber, M.E., and Zieg, G.A., 2000, Sulfur geo-
chemical constraints on Mesoproterozoic restricted marine deposition:
Lower Belt Supergroup, northwestern United States: Geochimica et Cos-
mochimica Acta, v. 64, p. 427437.
Lyons, T.W., Kah, L.C., and Gellatly, A.M., 2004, The Precambrian sulphur
isotope record of evolving atmospheric oxygen, in Eriksson, P.G., et al.,
eds., The Precambrian Earth: Tempos and events: Developments in Pre-
cambrian geology: Amsterdam, Elsevier, p. 421440.
Ma, G.L., Beaudoin, G., Qi, S J., and Li, Y., 2004: Geology and geochemistry
of the Changba SEDEX Pb-Zn deposit, Qinling orogenic belt, China: Min-
eralium Deposita, v. 39, p. 380395.
Ma, G.L., Beaudoin, G., Zhong, S.J., Li, Y., and Zeng, Z.R., 2007: Geology
and geochemistry of the Dengjiashan Zn-Pb SEDEX deposit, Qinling belt,
China: Canadian Journal of Earth Sciences, v. 44, p. 479492.
Melezhik, V.A., Lindahl, I., Pokrovsky, B., and Nilson, L.P., 2000, Sulphur
source and genesis of polymetallic sulphide occurrences of the Ofoten dis-
trict in the Central-North Norwegian Caledonides: Evidence from sulphur
isotopic studies: Mineralium Deposita, v. 35, p. 465489.
Melezhik, V.A., Fallick, A.E., Rychanchik, D.V., and Kuznetsov, A.B., 2005,
Palaeoproterozoic evaporites in Fennoscandia: Implications for seawater
sulphate, the rise of atmospheric oxygen and local amplification of the delta
C-13 excursion: Terra Nova, v. 17, p. 141148.
Meyer, K.M., and Kump, L.R., 2008, Oceanic euxinia in Earth history:
Causes and consequences: Annual Review of Earth and Planetary Sciences,
v. 36, p. 251288.
Mojzsis, S.J., Coath, C.D., Greenwood, J.P., McKeegan, K.D., and Harrison,
T.M., 2003, Mass-independent isotope effects in Archean (2.5 to 3.8 Ga)
sedimentary sulfides determined by ion microprobe analysis: Geochimica
et Cosmochimica Acta, v. 67, p. 16351658.
Monster, J., Appel, P.W.U., Thode, H.G., Schidlowski, M., Carmichael,
C.M., and Bridgwater, D., 1979, Sulfur isotope studies in Early Archaean
sediments from Isua, West Greenland - Implications for the antiquity of
bacterial sulfate reduction: Geochimica et Cosmochimica Acta, v. 43, p.
405413.
Muir, T.L., 2002, The Hemlo gold deposit, Ontario, Canada: Principal de-
posit characteristics and constraints on mineralization: Ore Geology Re-
views, v. 21, p. 166.
Nakai, S., Halliday, A.N., Kesler, S.E., Jones, H.D., Kyle, J.R., and Lane,
T.E., 1993, Rb-Sr dating of sphalerites from Mississippi Valley-type (MVT)
ore-deposits: Geochimica et Cosmochimica Acta, v. 57, p. 417427.
Ohmoto, H., Watanabe, Y., Ikemi, H., Poulson, S.R., and Taylor, B. E., 2006,
Sulphur isotope evidence for an oxic Archaean atmosphere: Nature, v. 442,
p. 908911.
Olcott, A.N., Sessions, A.L., Corsetti, F.A., Kaufman, A J., and de Oliviera,
T.F., 2005, Biomarker evidence for photosynthesis during Neoproterozoic
glaciation: Science, v. 310, p. 471474.
Olson, R.A., 1984, Genesis of paleokarst and strata-bound zinc-lead sulfide
deposits in a Proterozoic dolostone, northern Baffin Island, Canada: ECO-
NOMIC GEOLOGY, v. 79, p. 10561103.
Ono, S., 2008, Multiple-sulphur isotope biosignatures: Space Science Re-
views, v. 135, p. 203220.
Ono, S., Eigenbrode, J.L., Pavlov, A.A., Kharecha, P., Rumble, D., Kasting,
J.F., and Freeman, K.H., 2003, New insights into Archean sulfur cycle from
mass-independent sulfur isotope records from the Hamersley basin, Aus-
tralia: Earth and Planetary Science Letters, v. 213, p. 1530.
Ono, S., Wing, B., Johnston, D., Farquhar, J., and Rumble, D., 2006a, Mass-
dependent fractionation of quadruple stable sulfur isotope system as a new
tracer of sulfur biogeochemical cycles: Geochimica et Cosmochimica Acta,
v. 70, p. 22382252.
Ono, S., Beukes, N. J., Rumble, D., and Fogel, M. L., 2006b, Early evolution
of atmospheric oxygen from multiple-sulfur and carbon isotope records of
the 2.9 Ga Mozaan Group of the Pongola Supergroup, Southern Africa:
South African Journal of Geology, v. 109, p. 97108.
Ono, S., Shanks, W. C., Rouxel, O. J., and Rumble, D., 2007, S-33 constraints
on the seawater sulfate contribution in modern seafloor hydrothermal vent
sulfides: Geochimica et Cosmochimica Acta, v. 71, p. 11701182.
Ono, S.H., Beukes, N.J., and Rumble, D., 2009a, Origin of two distinct mul-
tiple-sulfur isotope compositions of pyrite in the 2.5 Ga Klein Naute For-
mation, Griqualand West basin, South Africa: Precambrian Research, v.
169, p. 4857.
Ono, S.H., Kaufman, A.J., Farquhar, J., Sumner, D.Y., and Beukes, N.J.,
2009b, Lithofacies control on multiple-sulfur isotope records and
Neoarchean sulfur cycles: Precambrian Research, v. 169, p. 5867.
Painter, M.G.M., Golding, S.D., Hannan, K.W., and Neudert, M.K., 1999,
Sedimentologic, petrographic and sulfur isotope constraints on fine-
grained pyrite formation at Mount Isa mine and environs, Northwest
Queensland, Australia: ECONOMIC GEOLOGY, v. 94, p. 883912.
Pandalai, H.S., Changkakoti, A., Krouse, H.R., and Gunalan, N., 1991, The
relationship between carbon, sulfur and pyritic iron in the Amjhore de-
posit, Bihar, India: ECONOMIC GEOLOGY, v. 86, p. 862869.
Papineau, D., Mojzsis, S.J., Coath, C.D., Karhu, J.A., and McKeegan, K.D.,
2005, Multiple sulfur isotopes of sulfides from sediments in the aftermath
of Paleoproterozoic glaciations: Geochimica et Cosmochimica Acta, v. 69,
p. 50335060.
Papineau, D., Mojzsis, S.J., and Schmitt, A.K., 2007, Multiple sulfur isotopes
from Paleoproterozoic Huronian interglacial sediments and the rise of at-
mospheric oxygen: Earth and Planetary Science Letters, v. 255, p. 188212.
530 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 530
Paradis, S., and Nelson, J.L., 2007, Metallogeny of the Robb Lake carbonate-
hosted zinc-lead district, northeastern British Columbia: in Goodfellow,
W.D., ed., Mineral deposits of Canada: A synthesis of major deposit-types,
district metallogeny, the evolution of geological provinces, and exploration
methods: Geological Association of Canada, Mineral Deposits Division,
Special Publication no. 5, p. 633654.
Partridge, M.A., Golding, S.D., Baublys, K.A., and Young, E., 2008, Pyrite
paragenesis and multiple sulfur isotope distribution in late Archean and
early Paleoproterozoic Hamersley basin sediments: Earth and Planetary
Science Letters, v. 272, p. 4149.
Paytan, A., Kastner, M., Campbell, D., and Thiemens, M.H., 1998, Sulfur
isotopic composition of Cenozoic seawater sulfate: Science, v. 282, p.
14591462.
2004, Seawater sulfur isotope fluctuations in the Cretaceous: Science, v.
304, p. 16631665.
Peevler, J., Fayek, M., Misra, K.C., and Riciputi, L.R., 2003, Sulfur isotope
microanalysis of sphalerite by SIMS: constraints on the genesis of Mis-
sissippi valley-type mineralization, from the Mascot-Jefferson City dis-
trict, East Tennessee: Journal of Geochemical Exploration, v. 80, p.
277296.
Perkins, C., 1996, Ar-40/Ar-39 age constraints on deformation and mineral-
ization, Rosebery Zn-Pb-Cu and Mount Lyell Cu deposits, Tasmania, Aus-
tralia: Mineralium Deposita, v. 31, p. 7183.
Peters, M., Strauss, H., Farquhar, J., Ockert, C., Eickmann, B., and Jost,
C.L., 2010, Sulfur cycling at the Mid-Atlantic Ridge: A multiple sulfur iso-
tope approach: Chemical Geology, v. 269, p. 180196.
Philippot, P., Van Zuilen, M., Lepot, K., Thomazo, C., Farquhar, J., and Van
Kranendonk, M. J., 2007, Early Archaean microorganisms preferred ele-
mental sulfur, not sulfate: Science, v. 317, p. 15341537.
Pope, M.C., and Grotzinger, J.P., 2003, Paleoproterozoic Stark Formation,
Athapuscow basin, northwest Canada: Record of cratonic-scale salinity cri-
sis: Journal of Sedimentary Research, v. 73, p. 280295.
Poulton, S., Bekker, A., and Canfield, D.E., 2009, Early Paleoproterozoic
fluctuations in biospheric oxygenation, Geochimica et Cosmochimica Acta,
v. 73, v. 13, p. A1047A1047
Poulton, S.W., Fralick, P.W., and Canfield, D.E., 2004, The transition to a sul-
phidic ocean similar to 1.84 billion years ago: Nature, v. 431, p. 173177.
Prokoph, A., Shields, G.A., and Veizer, J., 2008, Compilation and time-series
analysis of a marine carbonate delta O-18, delta C-13, Sr-87/Sr-86 and delta
S-34 database through Earth history: Earth-Science Reviews, v. 87, p.
113133.
Reinhard, C.T., Raiswell, R., Scott, C., Anbar, A.D., and Lyons, T.W., 2009,
A Late Archean sulfidic sea stimulated by early oxidative weathering of the
continents: Science, v. 326, p. 713716.
Rouxel, O.J., Bekker, A., and Edwards, K.J., 2005, Iron isotope constraints on
the Archean and Paleoproterozoic ocean redox state: Science, v. 307, p.
10881091.
Rye, D.M., and Williams, N., 1981, Studies of the base metal sulfide deposits
at McArthur River, Northern Territory, Australia. III: The stable isotope
geochemistry of the H.Y.C., Ridge, and Cooley deposits: ECONOMIC GEOL-
OGY, v. 76, p. 126.
Ryznar, G., Campbell, F.A., and Krouse, H.R., 1967, Sulfur isotopes and the
origin of the Quemont ore body, ECONOMIC GEOLOGY, v. 62, p. 664678.
Sangster, D.F., 1990, Mississippi Valley-Type and SEDEX lead zinc de-
positsa comparative-examination: Transactions of the Institution of Min-
ing and Metallurgy Section B-Applied Earth Science, v. 99, p. B21B42.
2002, The role of dense brines in the formation of vent-distal sedimen-
tary-exhalative (SEDEX) lead-zinc deposits: Field and laboratory evidence:
Mineralium Deposita, v. 37, p. 149157.
Sasaki, A., and Krouse, H.R., 1969, Sulfur isotopes and the Pine Point lead
zinc mineralization: ECONOMIC GEOLOGY, v. 64, p. 718730.
Schaefer, M.O., 2002, Paleoproterozoic Mississippi Valley-type Pb-Zn de-
posits of the Ghaap Group, Transvaal Supergroup in Griqualand West,
South Africa: Unpublished Ph.D. thesis, Johannesburg, South Africa, Rand
Afrikaans University, 373 p.
Schrder, S., Bekker, A., Beukes, N.J., Strauss, H., and van Niekerk, H.S.,
2008, Rise in seawater sulphate concentration associated with the Paleo-
proterozoic positive carbon isotope excursion: Evidence from sulphate
evaporites in the similar to 2.2-2.1 Gyr shallow-marine Lucknow Forma-
tion, South Africa: Terra Nova, v. 20, p. 108117.
Scott, C., Lyons, T.W., Bekker, A., Shen, Y., Poulton, S.W., Chu, X., an Anbar,
A.D., 2008, Tracing the stepwise oxygenation of the Proterozoic ocean: Na-
ture, v. 452, p. 456459.
Scott, K.M., Smith, J.W., Sun, S.-S., and Taylor, G.F., 1985, Proterozoic cop-
per deposits in NW Queensland, Australia: Sulfur isotopic data: Mineral-
ium Deposita, v. 20, p. 116126.
Shanks, W.C., Woodruff, L.G., Jilson, G.A., Jennings, D.S., Modene, J.S., and
Ryan, B.D., 1987, Sulfur and lead isotope studies of stratiform Zn-Pb-Ag
deposits, Anvil Range, Yukonbasinal brine exhalation and anoxic bottom-
water mixing: ECONOMIC GEOLOGY, v. 82, p. 600634.
Sharpe, R., and Gemmell, J.B., 2000, Sulfur isotope characteristics of the
Archean Cu-Zn Gossan Hill VHMS deposit, Western Australia: Mineral-
ium Deposita, v. 35, p. 533550.
Shen, Y.N., Canfield, D.E., and Knoll, A.H., 2002, Middle Proterozoic ocean
chemistry: Evidence from the McArthur Basin, northern Australia: Ameri-
can Journal of Science, v. 302, p. 81109.
Shen, Y., Knoll, A.H., and Walter, M.R., 2003, Evidence for low sulphate and
anoxia in a mid-Proterozoic marine basin: Nature, v. 423, p. 632635.
Shen, Y.N., Farquhar, J., Masterson, A., Kaufman, A.J., and Buick, R., 2009,
Evaluating the role of microbial sulfate reduction in the early Archean
using quadruple isotope systematics: Earth and Planetary Science Letters,
v. 279, p. 383391.
Slack, J.F., and Cannon, W.F., 2009, Extraterrestrial demise of banded iron
formations 1.85 billion years ago: Geology, v. 37, p. 10111014.
Slack, J.F., Dumoulin, J.A., Schmidt, J.M., Young, L.E., and Rombach, C.S.,
2004, Paleozoic sedimentary rocks in the Red Dog Zn-Pb-Ag district and
vicinity, western Brooks Range, Alaska: Provenance, deposition, and metal-
logenic significance: ECONOMIC GEOLOGY, v. 99, p. 13851414.
Slack, J.F., Grenne, T., Bekker, A., Rouxel, O.J., and Lindberg, P.A., 2007, Sub-
oxic deep seawater in the late Paleoproterozoic: Evidence from hematitic
chert and iron formation related to seafloor-hydrothermal sulfide deposits,
central Arizona, USA: Earth and Planetary Science Letters, v. 255, p. 243256.
Smith, J.W., Burns, M.S., and Croxford, N.J.W., 1978, Stable isotope studies
of the origins of mineralization at Mount Isa. I: Mineralium Deposita, v. 13,
p. 369381.
Solomon, M., Rafter, T.A., and Jensen M.L., 1969, Isotope studies on the
Rosebery, Mount Farrell and Mount Lyell ores, Tasmania: Mineralium De-
posita, v. 4, p. 172199.
Solomon, M., Eastoe, C. J., Walshe J.L., and Green G.R., 1988, Mineral de-
posits and sulfur isotope abundances in the Mount Read Volcanics between
Que River and Mount Darwin, Tasmania: ECONOMIC GEOLOGY, v. 83, p.
13071328.
Srensen, K.B., and Canfield, D.E., 2004, Annual fluctuations in sulfur iso-
tope fractionation in the water column of a euxinic marine basin: Geochim-
ica et Cosmochimica Acta, v. 68, p. 503515.
St. Marie, J., and Kesler, S.E., 2000, Iron-rich and iron-poor Mississippi Val-
ley-type mineralization, Metaline district, Washington: ECONOMIC GEOL-
OGY, v. 95, p. 10911106.
Stumm, W., and Morgan, J.J., 1981, Aquatic chemistry. An introduction em-
phasizing chemical equilibria in natural waters [2nd ed.]: New York, Wiley,
780 p.
Sverjensky, D.A., 1981, The origin of a Mississippi Valley-type deposit in the
Viburnum Trend, southeast Missouri: ECONOMIC GEOLOGY, v. 76, p. 1848
1872.
Taylor, B.E., and South B.C., 1985, Regional stable isotope systematics of hy-
drothermal alteration and massive sulfide deposition in the West Shasta
district, California: ECONOMIC GEOLOGY, v. 80, p. 21492163.
Thode, H.G., Ding, T., and Crocket, J.H., 1991, Sulfur-isotope and elemen-
tal geochemistry studies of the Hemlo gold mineralization, Ontario
sources of sulfur and implications for the mineralization process: Canadian
Journal of Earth Sciences, v. 28, p. 1325.
Thomazo, C., Ader, M., Farquhar, J., and Philippot, P., 2009, Methanotrophs
regulated atmospheric sulfur isotope anomalies during the Mesoarchean
(Tumbiana Formation, Western Australia): Earth and Planetary Science
Letters, v. 279, p. 6575.
Tornos, F., Clavijo, E.G., and Spiro, B., 1998, The Filon Norte orebody
(Tharsis, Iberian pyrite belt): A proximal low-temperature shale-hosted
massive sulphide in a thin-skinned tectonic belt: Mineralium Deposita, v.
33, p. 150169.
Tornos, F., Solomon, M., Conde, C., and Spiro, B.F., 2008, Formation of the
Tharsis massive sulfide deposit, Iberian pyrite belt: Geological, lithogeo-
chemical, and stable isotope evidence for deposition in a brine pool: ECO-
NOMIC GEOLOGY, v. 103, p. 185214.
Tupper, W.M., 1960, Sulfur isotopes and the origin of the sulfide deposits of
the Bathurst-Newcastle area of northern New Brunswick: ECONOMIC GE-
OLOGY, v. 55, p. 16761707.
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 531
0361-0128/98/000/000-00 $6.00 531
Turchyn, A.V., and Schrag, D.P., 2004, Oxygen isotope constraints on the sul-
fur cycle over the past 10 million years: Science, v. 303, p. 20042007.
2006, Cenozoic evolution of the sulfur cycle: Insight from oxygen iso-
topes in marine sulfate: Earth and Planetary Science Letters, v. 241, p.
763779.
Turner, R.J.W., 1992, Formation of Phanerozoic stratiform sediment-hosted
zinc lead depositsevidence for the critical role of ocean anoxia: Chemical
Geology, v. 99, p. 165188.
Turro, N.J., 1983, Influence of nuclear-spin on chemical-reactionsmag-
netic isotope and magnetic-field effects (a review): Proceedings of the Na-
tional Academy of Sciences of the United States of America-Physical Sci-
ences, v. 80, p. 609621.
Ueno, Y., Ono, S., Rumble, D., and Maruyama, S., 2008, Quadruple sulfur
isotope analysis of ca. 3.5 Ga Dresser Formation: New evidence for micro-
bial sulfate reduction in the early Archean: Geochimica et Cosmochimica
Acta, v. 72, p. 56755691.
Vanderwood, T.B., and Thiemens, M.H., 1980, The Fate of the hydroxyl rad-
ical in the earths primitive atmosphere and implications for the production
of molecular-oxygen: Journal of Geophysical Research-Oceans and Atmos-
pheres, v. 85, p. 16051610.
Vearncombe, S., Barley, M.E., Groves, D.I., McNaughton, N.J., Mikucki,
E.J., and Vearncombe, J.R., 1995, 3.26 Ga Black smoker-type mineraliza-
tion in the Strelley belt, Pilbara craton, Western-Australia: Journal of the
Geological Society, v. 152, p. 587590.
Velasco, F., Sanchez-Espana, J., Boyce, A.J., Fallick, A.E., Saez, R., and
Almodovar, G.R., 1998, A new sulphur isotopic study of some Iberian
pyrite belt deposits: evidence of a textural control on sulphur isotope com-
position: Mineralium Deposita, v. 34, p. 418.
Velasco, F., Herrero, J.M., Yusta, I., Alonso, J.A., Seebold, I., and Leach,
D.L., 2003: Geology and geochemistry of the Reocin zinc-lead deposit,
Basque-Cantabrian basin, northern Spain: ECONOMIC GEOLOGY, v. 98, p.
13711396.
Wagner, T., Boyce, A.J., and Fallick, A.E., 2002, Laser combustion analysis of
delta S-34 of sulfosalt minerals: Determination of the fractionation system-
atics and some crystal-chemical considerations: Geochimica et Cos-
mochimica Acta, v. 66, p. 28552863.
Walker, J.C.G., and Brimblecombe, P., 1985, Iron and sulfur in the pre-bio-
logic ocean: Precambrian Research, v. 28, p. 205222.
Walker, R.N., Gulson, B., and Smith, J., 1983, The Coxco deposit; a Protero-
zoic Mississippi Valley-type deposit in the McArthur River district, North-
ern Territory, Australia: ECONOMIC GEOLOGY, v. 78, p. 214249.
Watanabe, Y., Farquhar, J., and Ohmoto, H., 2009, Anomalous fractionations
of sulfur isotopes during thermochemical sulfate reduction: Science, v. 324,
p. 370373.
Whelan, J.F., Rye, R.O., Delorraine, W., and Ohmoto, H., 1990, Isotopic geo-
chemistry of a mid-Proterozoic evaporite basinBalmat, New York: Amer-
ican Journal of Science, v. 290, p. 396424.
Wilkinson, J.J., Eyre, S.L., and Boyce, A.J., 2005, Ore forming processes in
Irish-type carbonate-hosted Zn-Pb deposits: Evidence from mineralogy,
chemistry, and isotopic composition of sulfides at the Lisheen mine: ECO-
NOMIC GEOLOGY, v. 100, p. 6386.
Willan, R.C.R., and Coleman, M.L., 1983, Sulfur isotope study of the Aber-
feldy barite, zinc, lead deposit, and minor sulfide mineralization in the Dal-
radian metamorphic terrain, Scotland: ECONOMIC GEOLOGY, v. 78, p.
16191656.
Wille, M., Kramers, J.D., Nagler, T.F., Beukes, N.J., Schrder, S., Meisel, T.,
Lacassie, J.P., and Voegelin, A.R., 2006, Evidence for a gradual rise of oxy-
gen between 2.6 and 2.5 Ga from Mo isotopes and Re-PGE signatures:
Geochimica et Cosmochimica Acta, v. 71, p. 24172435.
Wu, N.P., Farquhar J., Strauss, H., Kim, S.T., and Canfield, D.E., 2010, Eval-
uating the S-isotope fractionation associated with Phanerozoic pyrite bur-
ial: Geochimica et Cosmochimica Acta, doi:10.1016/j.gca.2009.12.012.
Yakushev, E.V., and Neretin, L.N., 1997, One-dimensional modeling of ni-
trogen and sulfur cycles in the aphotic zones of the Black and Arabian Seas:
Global Biogeochemical Cycles, v. 11, p. 401414.
Yang, J., Large, R.R., and Bull, S.W., 2004, Factors controlling free thermal
convection in faults in sedimentary basins: implications for the formation of
zinc-lead mineral deposits: Geofluids, v. 4, p. 237247.
Zmolek, P., Xu, X.P., Jackson, T., Thiemens, M.H., and Trogler, W C., 1999,
Large mass independent sulfur isotope fractionations during the pho-
topolymerization of (CS2)-C-12 and (CS2)-C-13: Journal of Physical
Chemistry A, v. 103, p. 24772480.
532 FARQUHAR ET AL.
0361-0128/98/000/000-00 $6.00 532
The foundation of isotopic geochemistry is the existence of
processes that discriminate between isotopes. For sulfur
which has four stable isotopes, the host of processes that dis-
criminate between
34
S and
32
S do not discriminate between
33
S,
36
S, and
32
S in exactly the same way. This allows one to use
variations in the relative abundance of
32
S,
33
S,
34
S, and
36
S to
provide insight into changes in the way that sulfur is cycled in
geological environments. Sulfur isotope notation used here
includes
34
S,
33
S, and
36
S. The
34
S describes the devia-
tion of a measured
34
S/
32
S from a reference
34
S/
32
S, and the

33
S and
36
S describe the deviation of a measured
33
S/
32
S or
a measured
36
S/
32
S from reference relationship between
34
S/
32
S,
33
S/
32
S, and
36
S/
32
S. We choose these reference rela-
tionships to approximate single-step equilibrium isotope and
draw on an exponential relationship between fractionation
factors (e.g. (Craig, 1957), where is an
exponential term that is related to the relative mass differ-
ences between the different sulfur isotopes). We illustrate
this principle using thermodynamics of isotope exchange in-
volving elemental sulfur and hydrogen sulfide using an equa-
tion that describes the distribution coefficients:
, (1)
G
3k_32
S
is the Gibbs free energy associated with the ex-
change reaction involving species with
32
S and
3k
S (in this case
k is 3, 4, or 6), R is the natural gas constant, T is temperature
(in Kelvins), and the activities are given by a
species
. Because
isotope exchange reactions are ideal with respect to mixing
(Ono et al., 2006b; Farquhar et al., 2007b), activity equals
concentration, and we can rewrite these as:
, (2)
which is equivalent to: G
3k_32
S
= RT ln (
3k

SH
2
S
), and can be
rearranged to give , our basis for
defining
33
S in terms of an exponential relationship.
The values of G for the different isotope pairs arise from
differences in vibrational energies which are related to the
masses of the atoms. This is the reason these effects are re-
ferred to as mass-dependent. Other physical processes also
discriminate on the basis of mass (e.g., gravitational separa-
tion, which depends on principles of buoyancy) and are mass-
dependent, too, but are not bound by the same exponential
relationship between the fractionation factors () for the dif-
ferent isotopes. This characteristic can be used to identify
these processes in nature. Kinetic isotope effects arise as a re-
sult of differences in the rates of chemical reactions of species
with different isotopic composition. While the same type of
vibrational energy terms play a part in these effects, other fac-
tors can also play a role in determining the relative reaction
rates of different isotopic species such as: (1) terms that de-
scribe the energy related to the configuration of transition
state which is also treated as a vibrational term and referred
to as an imaginary vibrational mode that describes vibrations
in the reaction coordinate (Note that for some reactions, the
number of imaginary vibrational modes may depend on the
symmetry of the reacting or transition state species
Hathorne and Marcus, 2000; Gao and Marcus, 2001), (2)
terms that describe specific other controls related to elec-
tronic spin (MIE) (Turro, 1983; Buchachenko, 1995) or (3)
terms that relate to the matching of vibrational levels or po-
tential energy surfaces for bound and unbound states along
the reaction pathway (Zmolek et al., 1999; Bhattacharya et al.,
2000) or even (4) terms that describe the presence or absence
of vibrational energy levels in weakly bound states (Lasaga et
al., 2008). When kinetic isotope effects are strongly influ-
enced by factors other than the mass of the isotopes involved
in the reactions, they can produce anomalous and large
changes in
33
S and
36
S for small changes in
34
S, and be-
cause these are produced by factors other than mass, they
have been referred to as mass-independent or nonmass-de-
pendent isotope effects (MIF).
( ) ( )


36 33
1
36
1
33 34
= =
S H S
S H S
S
k
k
k
a a
a a
RT G
3
2
32
32
2
3
32 _ 3
ln

=
[ ] [ ]
[ ] [ ] S H S
S H S
RT G
k
k
S
k
3
2
32
32
2
3
ln
32 _ 3
=
( )
S
k
S
G
G
S H S
k
S H S
32 _ 3
32 _ 34
2 2
3 34


=
SULFUR CYCLE EVOLUTION, SULFUR ISOTOPES, SEDIMENTS, AND BASE METAL SULFIDE DEPOSITS 533
0361-0128/98/000/000-00 $6.00 533
APPENDIX 1.
Mass-Dependent and Mass-Independent Sulfur Isotope Geochemistry

You might also like