Sunteți pe pagina 1din 10

Sedimentary Geology 131 (2000) 7786

www.elsevier.nl/locate/sedgeo

Early diagenetic pyrite morphology in a mudstone-dominated


succession: the Lower Jurassic Cleveland Ironstone Formation,
eastern England
K.G. Taylor a,*, J.H.S. Macquaker b
a

Department of Environmental and Geographical Sciences, Manchester Metropolitan University, Manchester M1 5GD, UK
b
Department of Earth Sciences, University of Manchester, Manchester M13 9PL, UK
Received 29 March 1999; accepted 19 November 1999

Abstract
Diagenetic pyrite in the mudstones and ironstones of the Lower Jurassic Cleveland Ironstone Formation of eastern England
exhibits two distinct morphologies: framboidal pyrite, commonly associated with organic matter, and euhedral pyrite,
associated with detrital clay pellets. These two morphologies are mutually exclusive in occurrence. Framboidal pyrite is present
in clay-rich mudstones, ooidal ironstones, apatite-rich units and some silt-rich mudstones. Euhedral pyrite is present in silt-rich
and sand-rich mudstones. d 34S isotopic analysis of six samples of pyrite suggests that both types of pyrite morphology
precipitated during early diagenesis from porewaters with open access to overlying sea-water, although both probably acted
as sites for continued pyrite precipitation during burial. It is proposed that framboidal pyrite precipitated from iron-dominated
porewaters at sites of sulfide supply (i.e. in the region of organic matter as a result of bacterial sulfate reduction) where, locally,
sulfide production rates were high enough for porewaters to reach supersaturation with respect to FeS. Euhedral pyrite also
precipitated from iron-dominated porewaters, but sulfide production rates from organic matter was such that FeS saturation was
not reached at the sites of sulfide production. Instead, euhedral pyrite was precipitated directly from porewater when FeS2
saturation was reached. The control over pyrite morphology was probably the amount and reactivity of the organic matter
within the deposited sediments. The sand-rich mudstones contained less reactive organic matter due to clastic dilution and
deposition in shallower environments with O2-rich bottom waters. The ironstones and apatite-rich units were deposited under
very low sedimentation rates, and as a result organic matter contents were very low and iron reduction dominated early
diagenesis, which inhibited sulfate-reduction. The presence of minor framboidal pyrite within these units, however, suggests
that sulfide reduction took place in micro-environments during early diagenesis. q 2000 Elsevier Science B.V. All rights
reserved.
Keywords: pyrite; diagenesis; mudstone; ironstone; Jurassic; England

1. Introduction
Pyrite (FeS2) is a common mineral product of early
diagenesis in organic-rich sediments. It results from
* Corresponding author. Fax: 1 44-161-247-6318.
E-mail address: k.g.taylor@mmu.ac.uk (K.G. Taylor).

the reaction of sulfide (produced via bacterial sulfate


reduction, Berner, 1970) with either Fe(III) in
sediments or Fe(II) produced by bacterial Fe(III)
reduction (Lovley, 1991). Recent work on iron, sulfur
and carbon geochemistry in sediments and, in particular, the nature and mechanism of early diagenetic
pyrite formation, has led to the development of

0037-0738/00/$ - see front matter q 2000 Elsevier Science B.V. All rights reserved.
PII: S0037-073 8(00)00002-6

78
K.G. Taylor, J.H.S. Macquaker / Sedimentary Geology 131 (2000) 7786
Fig. 1. Sedimentological and stratigraphic log of the Cleveland Ironstone Formation, showing the stratigraphic distribution of lithofacies types and the location of framboidal and
euhedral pyrite dominated samples. Sample locations used in this study are indicated by dots on the side of the column. Only those samples that are shown in Table 1 are labeled. For
a full listing of the sample set see Macquaker and Taylor (1996).

K.G. Taylor, J.H.S. Macquaker / Sedimentary Geology 131 (2000) 7786

geochemical paleoenvironmental indicators. Of


particular note is the application of carbon to sulfur
ratios within sediments (Berner and Raiswell, 1984;
Leventhal, 1995) and the degree of pyritization (DOP;
Raiswell et al., 1988). More recently, textural features
of early diagenetic pyrite have been used to infer
geochemical and environmental conditions. Passier
et al. (1997) showed that pyrite morphology within
a marine sapropel and associated sediments could
provide information on early diagenetic processes,
whilst Wilkin et al. (1996) argued that the size of
pyrite framboids could help distinguish oxic and
anoxic environments.
Early diagenetic pyrite can have varied morphologies. Two of the most commonly observed
morphologies are framboids (spherical aggregates of
micron-sized pyrite crystals; Love and Amstutz,
1966; Sweeney and Kaplan, 1973) and micron-sized
euhedral crystals (Raiswell, 1982; Passier et al.,
1997). Models have been put forward that relate pyrite
morphology to saturation levels of Fe(II) and sulfide
in sediment porewaters, and, therefore, to the rate of
production of Fe and sulfide during early diagenesis
(Raiswell, 1982; Passier et al., 1997). It has been
argued that euhedral pyrite precipitates from porewaters oversaturated with respect to pyrite but undersaturated with respect to iron monosulfides
(Goldhaber and Kaplan, 1974; Raiswell, 1982;
Wang and Morse, 1996; Rickard, 1997). In contrast,
pyrite framboids are commonly believed to have
precipitated, via iron monosulfide intermediates,
from porewaters supersaturated with respect to both
pyrite and iron monosulfides (Sweeney and Kaplan,
1973; Morse et al., 1987; Roberts and Turner, 1993).
Studies of iron, sulfur and carbon geochemistry of
modern sediments has led to an improved understanding of the controls on pyrite formation within marine
sediments (Canfield, 1989; Canfield et al., 1992; Raiswell and Canfield, 1996; Raiswell, 1997). A key control
on pyrite formation is the amount and reactivity of
organic matter within the sediment, which controls
the rate at which sulfide is produced by sulfate-reducing
bacteria (Berner, 1970; Westrich and Berner, 1984). In
addition, the nature of iron and sulfide reactions within
sediment porewaters is dependent on the reactivity of
detrital iron minerals within the sediment (Canfield et
al., 1992; Raiswell, 1997). The nature and style of pyrite
formation is dependent on whether porewaters are

79

21

dominated by Fe or sulfide (Canfield and Raiswell,


1991; Raiswell, 1997). In the majority of marine sediments, during earliest diagenesis porewater sulfide is
buffered to low levels by highly reactive iron oxides,
resulting in iron-dominated porewaters (dissolved iron
concentrations q dissolved sulfide). Depending on the
rate of sulfide production, pyrite will form either at or
away from sites of sulfide production. After consumption of reactive iron oxides, sulfide builds up in
porewaters, commonly leading to a deeper zone of
sulfide-dominated porewaters (Canfield and Raiswell,
1991). These models have also led to a better understanding of fossilization processes, whereby the nature
of early diagenetic porewaters (Fe 21- or sulfidedominated) has a major control on soft body pyritization
(Canfield and Raiswell, 1991; Raiswell, 1997).
Detailed observation of early diagenetic pyrite form
and location in mudstone-dominated sedimentary
successions are limited. In this paper we show that
within an ancient shallow marine mudstone succession (the Lower Jurassic Cleveland Ironstone
Formation of eastern England), early diagenetic pyrite
morphology varies in a systematic manner. In some
parts of the succession framboidal pyrite is the dominant morphology of early diagenetic pyrite, whereas
in other parts of the succession euhedral pyrite is the
dominant morphology of early diagenetic pyrite. On
the basis of integrated sedimentological and geochemical data we argue that the ultimate controls on
early diagenetic pyrite morphology in this succession
were sedimentary in nature (sedimentation rate,
organic matter content, bottom-water oxygenation),
which exerted a major control on carbon, sulfur and
iron reactivities.

2. The Cleveland Ironstone Formation, eastern UK


The Lower Jurassic (Upper Liassic) Cleveland
Ironstone Formation is well exposed at Staithes on
the Yorkshire Coast, in northeastern England
(Fig. 1). The succession comprises mudstones, ooidal
ironstones and phosphatic concretionary horizons
(Fig. 1; Rawson et al., 1982; Macquaker and Taylor,
1996). Macquaker and Taylor (1996) studied the
succession through a combination of field and thin
section analysis and recognized the presence of
clay-rich mudstones (.80% clay, ,10% silt 1 sand),

80

K.G. Taylor, J.H.S. Macquaker / Sedimentary Geology 131 (2000) 7786

Table 1
Pyrite d 34S, pyrite-S, total Fe, HCl-soluble Fe, total organic carbon and DOP values for selected samples containing framboidal or euhedral
pyrite
Sample

d 34S ( CDT)

Framboidal pyrite
ST-11
232.5
ST-13
14.0
ST-14
28.7
Euhedral pyrite
ST-20
228.2
ST-21
211.9
ST-22
28.7

Pyrite-S (wt%)

Total Fe (wt%)

HCl-soluble Fe (wt%)

DOP

TOC (wt%)

0.87
1.63
1.19

6.10
4.45
6.95

2.15
1.80
1.92

0.25
0.43
0.34

0.70
0.88
0.90

0.40
1.30
0.67

5.25
5.07
5.75

2.31
2.15
2.30

0.13
0.32
0.19

0.94
0.76
0.74

silt-rich mudstones (.80% clay, 1020% silt 1


sand), sand-rich mudstones (.50% clay, .20% silt 1
sand), ooidal ironstones (predominantly berthierine
ooids and early diagenetic pore-filling berthierine
and siderite cement) and apatite-rich units (predominantly early diagenetic apatite and berthierine
cements) (Fig. 2). In addition, Macquaker and Taylor
(1996, 1997) recognized the presence of systematic
variations in detrital grain size within the mudstones,
with small-scale (0.0010.05 m), intermediate-scale
(0.11 m) and large-scale (13 m) upward-coarsening units and upward-fining units present. These
were interpreted to represent beds, parasequences
and systems tracts, respectively (sensu Campbell,
1967; Van Wagoner et al., 1990) formed in response
to variations in relative sea-level and accommodation
within the basin (Macquaker and Taylor, 1996). Clayrich mudstones and apatite-rich units were deposited
during times of increased accommodation and
reduced sedimentation rates, with apatite-rich units
forming at times of negligible sedimentation rate.
Sand-rich mudstones were deposited at higher
sedimentation rates during times of accommodation
decrease and shortening of the sediment transport
path. Ooidal ironstones were deposited during times
of sediment bypass (shallow water with extensive
sediment reworking) during times of lowered sealevel and negligible accommodation (Taylor and
Curtis, 1995; Macquaker and Taylor, 1996).

3. Material and methods


Samples were collected from coastal exposures at
Staithes, North Yorkshire, with the freshest, most

recently exposed sections being sampled. Samples


from the best-characterized cycle within the succession (that from the Avicula Ironstone seam to the
Raisdale Ironstone seam; Macquaker and Taylor,
1996) were chosen for this study (Fig. 1). Total iron
contents of samples were determined by the wet
chemical method of Aller et al. (1986). Samples
were heated in a muffle furnace at 4508C for 12 h
followed by digestion in hot 12 M HCl for 12 h.
Iron potentially reactive toward sulfide was determined by digestion in cold 1 M HCl. Samples were
digested at room temperature for 24 h. It is felt that
this method slightly underestimates iron potentially
reactive to sulfide, but has the benefit that it is less
prone to analytical error (Raiswell et al., 1994). All
iron concentrations of resulting solutions were
measured by atomic absorption spectrometry. Total
organic carbon contents in the mudstone samples
were determined by oxidation with potassium
dichromate, followed by titration against ammonium
iron (II) sulfate.
Polished thin sections 20 mm thick were made of all
samples, with the grinding and polishing being carried
out in oil to avoid sample disintegration (see
Macquaker and Gawthorpe, 1993). After initial
optical petrographic work in transmitted light, the
sections were coated with carbon and analyzed both
petrographically and chemically in a JEOL 6400
scanning electron microscope equipped with a
4-quadrant Link backscattered electron detector and
fully quantitative Link eXL energy dispersive X-ray
microanalysis system.
Six mudstone samples were selected for pyrite d 34S
analysis at the University of Leeds; three containing
framboidal pyrite and three containing euhedral

K.G. Taylor, J.H.S. Macquaker / Sedimentary Geology 131 (2000) 7786

81

Fig. 2. Lithofacies within the Cleveland Ironstone Formation. (A) Backscattered electron image (BSEI) of a clay-rich mudstone. Note the
presence of silt-sized quartz grains (Q) within a detrital clay matrix. Scale bar 10 mm: (B) BSEI of a silt-rich mudstone. Note the increase in
silt-sized quartz grains (Q) compared to the clay-rich mudstone. Scale bar 10 mm: (C) BSEI of a sand-rich mudstone. Silt and sand-sized
quartz grains (Q), with detrital clay grains and matrix. Scale bar 10 mm: (D) Optical micrograph of ooidal ironstone. Note the presence of
berthierine ooids (b) with concentric cortices, and pore-filling siderite cement (s). Scale bar 50 mm: (E) BSEI of apatite-rich unit. Rare quartz
grains (Q) are cemented by apatite and berthierine cement (bright areas). Very bright rims are edge-effects. Scale bar 10 mm:

82

K.G. Taylor, J.H.S. Macquaker / Sedimentary Geology 131 (2000) 7786

pyrite. Pyrite contents within the ooidal ironstones


and apatite-rich units are minor (p0.5%) and, as a
result, pyrite d 34S analysis was not undertaken on
these units. Pyrite sulfur was extracted using
chromous chloride as described by Newton et al.
(1995) and sulfide recovered as CuS for isotopic
analysis. Sulfide yield was determined by titration of
the remaining copper in the trap. CuS precipitates
were combusted with Cu2O under vacuum at
11708C to produce SO2 for isotopic analysis
(Robinson and Kusakabe, 1975). Sample gases were
purified by standard vacuum-cryogenic techniques
and analyzed on a VG SIRA10 triple collector, gas
source, isotope ratio mass spectrometer. Raw data
were corrected by standard procedures (e.g. Craig,
1957) and results reported in standard delta notation
in per mille () relative to the Canyon Diablo Troilite
(CDT) standard. Errors, estimated from replicate
analysis of standardized materials, are ^0.2 for S.
The chromous chloride pyrite extraction procedure
carries a small isotopic fractionation (Newton et al.,
1995) for which the data have been corrected.
For these samples only, pyrite Fe was calculated
from the pyrite S data, and DOP was determined
DOP pyrite Fe=pyrite Fe 1 HCl soluble iron;
Raiswell et al., 1988).

4. Results
Two morphologies of diagenetic pyrite are present
within the Cleveland Ironstone Formation cycle
studied: framboidal pyrite and euhedral pyrite.
These two pyrite morphologies are mutually exclusive
in virtually all samples studied, with each having a
distinct distribution throughout the succession. Within
the apatite-rich units, clay-rich mudstones, ooidal
ironstones and some silt-rich mudstones, pyrite occurs
as framboids (Fig. 1). These framboids are 15 mm in
size (Fig. 3(A)(C)), and are often associated with
organic matter (e.g. precipitated within organic matter
particles). In contrast, euhedral pyrite is present in
sand-rich mudstones and some silt-rich mudstones,
and is restricted to sediments up to 100 cm below
the Raisdale ooidal ironstone seam and to the top
part of the underlying parasequence (Fig. 1). Therefore, although either framboidal or euhedral pyrite can
occur within silt-rich mudstones, euhedral pyrite

occurs in those silt-rich mudstones in the upper,


shallower parts of parasequences. These euhedral
pyrite crystals are small (,1 mm) and are seen in
association with either clay mineral grains or pellets
of clay minerals (Fig. 3(D) and (E)). These euhedral
pyrite crystals occur around, and between, detrital
clay grains and sheets, but do not actually replace
clay minerals themselves. Throughout the Cleveland
Ironstone Formation there is evidence that earlier
crystals of diagenetic pyrite have acted as sites of
nucleation and overgrowth of later diagenetic pyrite
(e.g. framboidal pyrite grains cemented by pyrite).
Total Fe contents within the succession are 4.4
7.4 wt% in mudstone samples (16 samples) and
18.927.8 wt% in ooidal ironstones and apatite-rich
units (four samples). HCl-soluble Fe was 1.3
2.4 wt% for mudstones and 5.427.0 wt% for ironstone and apatite-rich units. There is no significant
difference in either total Fe or HCl-soluble Fe between
mudstones containing framboidal pyrite and
mudstones containing euhedral pyrite.
Six mudstone samples were analyzed for d 34S,
three samples containing framboidal pyrite and three
samples containing euhedral pyrite. d 34S data for
pyrite within these samples shows a large spread in
values from 232.5 CDT to 1 4.0 CDT (Table 1).
Although there is a large range in values, ranges for
each set are similar (although slightly larger for the
framboidal pyrite samples), with both sets of samples
possessing a minimum d 34S of around 230 CDT
(Table 1). Pyrite-S was only determined for these six
samples and DOP values range from 0.13 to 0.43
(Table 1). There is no significant difference in DOP
between framboidal pyrite and euhedral pyrite
samples, although samples containing euhedral pyrite
have the lowest values (Table 1). Total organic carbon
contents within the mudstone samples exhibit a small
range of values, from 0.6 to 0.9%, and no significant
differences were observed between samples containing framboidal pyrite and those containing euhedral
pyrite.

5. Discussion
The presence of two distinct forms of pyrite
morphology within different lithofacies in the
Cleveland Ironstone Formation succession suggests

K.G. Taylor, J.H.S. Macquaker / Sedimentary Geology 131 (2000) 7786

83

Fig. 3. Backscatter electron images of pyrite within the sediments of the Cleveland Ironstone Formation. (A) Cluster of framboidal pyrite (p) in a
clay-rich mudstone. Scale bar 10 mm: (B) Framboidal pyrite (p) in apatite-rich unit. Scale bar 10 mm: (C) Pyrite framboid (p) within an
ooidal ironstone sample. Note the presence of siderite cement (s) and calcite cement (c). Scale bar 10 mm: (D,E) Euhedral pyrite set within
clay grains and pellets within a sand-rich mudstone. Note the presence of small euhedral pyrite crystals (up to 1 mm in size) associated with
detrital clay minerals grains. Black areas in the images are organic matter. Scale bar 10 mm:

that early diagenetic porewaters and processes


differed in different parts of the succession. Raiswell
(1982) interpreted euhedral pyrite within Jurassic
concretions to have formed from porewaters at sulfide
saturation levels lower than that of iron monosulfides,
whilst pyrite framboids within the same concretions
formed rapidly from solutions highly supersaturated
with respect to both pyrite and iron monosulfides.
Raiswell (1982); Coleman and Raiswell (1995)
argued that framboidal pyrite formed during earliest
diagenesis, with euhedral pyrite forming later in burial
as dissolved sulfide concentrations decreased. This
model was supported by d 34S data, which indicated
that euhedral pyrite formed from a sulfate source
depleted in 32S, i.e. cut off from overlying sea-water
(Raiswell, 1982).

Petrographic and isotopic data, however, do not


support this scenario for the pyrite in the Cleveland
Ironstone Formation. Although both pyrite morphologies exhibit a wide range in d 34S, the lowest d 34S
value for each pyrite morphology is approximately
230 CDT (Table 1). This suggests that both pyrite
morphologies began to form from sulfate that was not
depleted in 32S, i.e. porewaters with open contact to
overlying seawater. Petrographic observations indicate that, in many samples, early diagenetic pyrite
has been overgrown by later pyrite. Therefore, the
spread in pyrite d 34S data most probably represents
a mixture of earlier pyrite with later pyrite. This is
supported by the general agreement between pyritesulfur contents and d 34S values (Table 1). Samples
with higher pyrite-sulfur contents have more

84

K.G. Taylor, J.H.S. Macquaker / Sedimentary Geology 131 (2000) 7786

S-depleted pyrite-d S values, suggesting a greater


proportion of later pyrite. Therefore, both pyrite
morphologies in the Cleveland Ironstone Formation
began to form during very early diagenesis. We, therefore, discount the possibility that framboidal pyrite
formed during earliest diagenesis, and euhedral pyrite
formed during later diagenesis. This assertion is
further supported by the fact that the two pyrite
morphologies are virtually always mutually exclusive
in their occurrence. The absence of pyrite framboids
in the euhedral pyrite-containing samples indicates
that there was not an earlier phase of framboidal pyrite
precipitation in these sediments.
There is no significant difference in total Fe and
HCl-soluble Fe between mudstones containing
framboidal pyrite and mudstones containing euhedral
pyrite. It would seem, therefore, unlikely that variations in iron contents were the cause of the differing
early diagenetic porewater chemistries between
framboidal pyrite and euhedral pyrite-containing
mudstones. DOP values for both suggest oxygenated
bottom-water conditions, and are consistent with data
for a range of modern and ancient sediments (Raiswell
et al., 1988; Raiswell and Canfield, 1998). Although
the data is limited (Table 1), sand-rich mudstone and
silt-rich mudstone samples containing euhedral pyrite
have slightly lower DOP values than samples containing framboidal pyrite, which may indicate more
oxygenated bottom-water conditions than for those
mudstones containing framboidal pyrite. Observed
increased levels of bioturbation within sand-rich
mudstones also support this interpretation that depositional bottom waters were more O2-rich than during
deposition of the clay-rich mudstones. Although no
significant difference was noted in TOC contents
between framboidal pyrite-containing and euhedral
pyrite-containing samples, this does not necessarily
imply that depositional organic matter contents and,
more importantly, organic matter reactivities were
similar. Such information is difficult to assess for
ancient sediment successions.
We propose that the control on the pyrite morphology within the Cleveland Ironstone Formation
mudstones was the rate of sulfide production during
early diagenesis, which determined whether FeS or
FeS2 saturation was reached in early diagenetic porewaters. In clay-rich and silt-rich mudstones, containing framboidal pyrite, sulfide production rates were
32

34

high as a result of low O2 contents in depositional


bottom waters and high organic matter reactivity,
leading to high bacterial sulfate reduction rates. As a
result porewater FeS saturation was reached locally,
leading to the ultimate formation of framboidal pyrite.
Within sand-rich mudstones, organic matter contents
were lower, due to both dilution by coarser detrital
material and pre-burial oxidation of organic matter in
O2-rich depositional bottom waters. As a result,
sulfide production rates were not high enough to
lead to FeS saturation. Instead, euhedral pyrite precipitated directly from porewaters as FeS2 saturation
was reached. Passier et al. (1997) have similarly
argued for a sulfide-production rate control on early
diagenetic framboidal and euhedral pyrite formation
in a sapropelic horizon. The presence of euhedral
pyrite in association with detrital clay pellets most
likely reflects the fact that reactive detrital iron within
the sediments was associated with these clay pellets.
The ooidal ironstones within the Cleveland
Ironstone Formation succession were deposited
within the shallowest waters within the succession
(Macquaker and Taylor, 1996). These sediments
therefore experienced extensive sediment reworking
and resuspension (Macquaker and Taylor, 1996).
Under such conditions, high iron oxide contents,
coupled with low organic matter contents leads to
the predominance of iron reduction during early
diagenesis and the precipitation of berthierine and
siderite cements (Aller et al., 1986; Taylor and Curtis,
1995). Apatite-rich units within the Cleveland
Ironstone Formation were deposited under conditions
of low net sediment accumulation (Macquaker and
Taylor, 1996). These low sediment accumulation
rates led to the burial of low contents of refractory
organic carbon, leading to the predominance of
suboxic diagenesis and berthierine and apatite precipitation. However, in both ooidal ironstone and
apatite-rich units the minor pyrite that is present is
in the form of framboids. This suggests that whilst
suboxic diagenesis dominated, locally high sulfate
production also took place within micro-environments during early diagenesis.

6. Conclusions
Early diagenetic pyrite morphology changed in

K.G. Taylor, J.H.S. Macquaker / Sedimentary Geology 131 (2000) 7786

response to depositional environment in the Jurassic


Cleveland Ironstone Formation. During deposition of
clay-rich mudstones and silt-rich mudstones at slow to
moderate sedimentation rates iron-dominated porewaters were prevalent during early diagenesis. High
sulfide production rates at sites of bacterial sulfate
reduction led to porewater supersaturation in FeS
and ultimate framboidal pyrite formation. During
deposition of silt-rich mudstones and sand-rich
mudstones under conditions of shallower water,
iron-dominated porewaters were also prevalent during
early diagenesis. However, lower organic matter
contents and reactivity led to lower rates of sulfide
production and porewater FeS saturation was not
reached, and euhedral pyrite precipitated directly.
Within ooidal ironstones and apatite-rich units only
very minor pyrite precipitated, as a result of suboxic
diagenesis dominating. However, scarce framboidal
pyrite grains within these units suggests that sulfate
reduction was active in microenvironments during
early diagenesis.
Such depositional controls on early diagenetic
pyrite morphology have not been detailed previously.
Although significant work on the style of fossil preservation through pyritization has been undertaken (see
Raiswell, 1997), systematic data on the form of early
diagenetic pyrite within mudstones, especially away
from concretions, is lacking. On the basis of the observations presented in this paper such data, coupled with
other petrographic observations, would prove of value
in interpreting the sedimentology and diagenesis of
mudstone-dominated successions.

Acknowledgements
Rob Raiswell and John Bloch are thanked for
constructive comments, which significantly improved
this paper. Simon Bottrell, University of Leeds is
thanked for carrying out the pyrite sulfur d 34S
analyses.

References
Aller, R.C., Mackin, J.E., Cox, R.T., 1986. Diagenesis of Fe and S in
Amazon inner shelf muds: apparent dominance of Fe reduction
and implications for the genesis of ironstones. Cont. Shelf Res.
6, 263289.

85

Berner, R.A., 1970. Sedimentary pyrite formation. Am. J. Sci. 268,


123.
Berner, R.A., Raiswell, R., 1984. C/S method for distinguishing
freshwater from marine sedimentary rocks. Geology 12, 365
368.
Campbell, C.V., 1967. Lamina, laminaset, bed and bedset.
Sedimentology 8, 726.
Canfield, D.E., 1989. Reactive iron in marine sediments. Geochim.
Cosmochim. Acta 53, 619632.
Canfield, D.E., Raiswell, R., 1991. Pyrite formation and fossil
preservation. In: Allison, P.S.A., Briggs, D.E.G. (Eds.).
Taphonomy: Releasing the Data Locked in the Fossil Record,
Plenum Press, New York, pp. 337387.
Canfield, D.E., Raiswell, R., Bottrell, S.H., 1992. The reactivity of
sedimentary iron minerals toward sulfide. Am. J. Sci. 292,
659683.
Coleman, M., Raiswell, R., 1995. Source of carbonate and origin of
zonation in pyritiferous carbonate concretions: evaluation of a
dynamic model. Am. J. Sci. 295, 282308.
Craig, H., 1957. Isotopic standards for carbon and oxygen and
correction factors for mass spectrometric analysis of carbon
dioxide. Geochim.Cosmochim. Acta 12, 133149.
Goldhaber, M.B., Kaplan, I.R., 1974. In: Goldberg, E.D. (Ed.). The
Sulfur Cycle, The Sea, vol. 5. Wiley, New York, pp. 569655.
Leventhal, J.S., 1995. Carbonsulfur plots to show diagenetic and
epigenetic sulfidation in sediments. Geochim. Cosmochim. Acta
59, 12071211.
Love, L.G., Amstutz, G.C., 1966. Review of microscopic pyrite
from the Devonian Chattanooga Shale and Rammelsberg
Banderz. Fortschr Mineralium 43, 272309.
Lovley, D.R., 1991. Dissimilatory Fe(III) and Mn(IV) reduction.
Microbiol. Rev. 55, 259287.
Macquaker, J.H.S., Gawthorpe, R.L., 1993. Mudstone lithofacies in
the Kimmeridge Clay Formation, Wessex Basin, southern
England: implications for the origin and controls of the distribution of mudstones. J. Sediment. Petrol. 63, 11291143.
Macquaker, J.H.S., Taylor, K.G., 1996. A sequence-stratigraphic
interpretation of a mudstone-dominated succession: the Lower
Jurassic Cleveland Ironstone Formation, UK. J. Geol. Soc.
London 153, 759770.
Macquaker, J.H.S., Taylor, K.G., 1997. Discussion on a sequence
stratigraphic interpretation of a mudstone-dominated succession: the Lower Jurassic Cleveland Ironstone Formation, UK.
J. Geol. Soc. London 154, 913916.
Morse, J.W., Millero, F.J., Cornwell, J.C., Rickard, D., 1987. The
chemistry of the hydrogen sulfide and iron sulfide systems in
natural waters. Earth Sci. Rev. 24, 142.
Newton, R.J., Bottrell, S.H., Dean, S.P., Hatfield, D., Raiswell, R.,
1995. An evaluation of the chromous chloride reduction method
for isotopic analyses of pyrite in rocks and sediment. Chem.
Geol. 125, 317320.
Passier, H.F., Middelburg, J.J., de Lange, G.J., Bottcher, M.E.,
1997. Pyrite contents, microtextures, and sulfur isotopes in
relation to formation of the youngest eastern Mediterranean
sapropel. Geology 25, 519522.
Raiswell, R., 1982. Pyrite texture, isotopic composition and the
availability of iron. Am. J. Sci. 82, 12441263.

86

K.G. Taylor, J.H.S. Macquaker / Sedimentary Geology 131 (2000) 7786

Raiswell, R., 1997. A geochemical framework for the application of


stable sulphur isotopes to fossil pyritisation. J. Geol. Soc.
London 154, 343356.
Raiswell, R., Canfield, D.E., 1996. The kinetics of reaction between
silicate iron and dissolved sulfide in Peru Margin sediments.
Geochim. Cosmochim. Acta 60, 27772787.
Raiswell, R., Canfield, D.E., 1998. Sources of iron for pyrite
formation in marine sediments. Am. J. Sci. 298, 219245.
Raiswell, R., Buckley, F., Berner, R.A., Anderson, T.F., 1988.
Degree of pyritisation of iron as a paleoenvironmental indicator
of bottom water oxygenation. J. Sediment. Petrol. 58, 812819.
Raiswell, R., Canfield, D.E., Berner, R.A., 1994. A comparison of
iron extraction methods for the determination of degree of
pyritisation and the recognition of iron-limited pyrite formation.
Chem. Geol. 111, 101110.
Rawson, P.F., Greensmith, J.T., Shalaby, S.E., 1982. Coarseningupwards cycles in the uppermost Staithes and Cleveland
Ironstone Formations (Lower Jurassic) of the Yorkshire Coast,
England. Proc. Geol. Assoc. 94, 9193.
Rickard, D., 1997. Kinetics of pyrite formation by the H2S oxidation
of iron (II) monosulfide in aqueous solutions between 25 and
1258C: the rate equation. Geochim. Cosmochim. Acta 61,
115134.
Roberts, A.P., Turner, G.M., 1993. Diagenetic formation of ferrimagnetic iron sulphide minerals in rapidly deposited marine

sediments, South Island, New Zealand. Earth Planet. Sci. Lett.


115, 257273.
Robinson, B.W., Kusakabe, M., 1975. Quantitative preparation of
sulfur dioxide for 32S/ 34S analysis by combustion with cuprous
oxide. Anal. Chem. 47, 11791181.
Sweeney, R.E., Kaplan, I.R., 1973. Pyrite framboid formation:
laboratory synthesis and marine sediments. Econ. Geol. 68,
618634.
Taylor, K.G., Curtis, C.D., 1995. Stability and facies association of
early diagenetic mineral assemblages: an example from a
Jurassic ironstonemudstone succession. J. Sediment. Res.
A65, 358368.
Van Wagoner, J.C., Mitchum, R.M., Campion, K.M., Rahmanian,
V.D., 1990. Siliciclastic sequence stratigraphy in well logs,
cores and outcrops: concepts for high-resolution correlation of
time and facies. AAPG Methods in Exploration Series 7, p. 55.
Wang, Q., Morse, J.W., 1996. Pyrite formation under conditions
approximating those in anoxic sediments I. Pathways and
morphology. Mar. Chem. 52, 99121.
Westrich, J.T., Berner, R.A., 1984. The role of sedimentary organic
matter in bacterial sulfate reduction: the G model tested.
Limnol. Oceanogr. 29, 236249.
Wilkin, R.T., Barnes, H.L., Brantley, S.L., 1996. The size distribution of framboidal pyrite in modern sediments: an indicator of
redox conditions. Geochim. Cosmochim. Acta 60, 38973912.

S-ar putea să vă placă și