You are on page 1of 34

Earth-Science Reviews 95 (2009) 63–96

Contents lists available at ScienceDirect

Earth-Science Reviews
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / e a r s c i r ev

A marine biogeochemical perspective on black shale deposition


D.Z. Piper a,⁎, S.E. Calvert b
a
US Geological Survey, M/S 901, Menlo Park CA., 94025, United States
b
University of British Columbia, Vancouver B.C., Canada V6T 1W5

a r t i c l e i n f o a b s t r a c t

Article history: Deposition of marine black shales has commonly been interpreted as having involved a high level of marine
Received 22 November 2007 phytoplankton production that promoted high settling rates of organic matter through the water column and
Accepted 11 March 2009 high burial fluxes on the seafloor or anoxic (sulfidic) water-column conditions that led to high levels of
Available online 28 March 2009
preservation of deposited organic matter, or a combination of the two processes. Here we review the
hydrography and the budgets of trace metals and phytoplankton nutrients in two modern marine basins that
Keywords:
black shales
have permanently anoxic bottom waters. This information is then used to hindcast the hydrography and
trace elements biogeochemical conditions of deposition of a black shale of Late Jurassic age (the Kimmeridge Clay
marine geochemistry Formation, Yorkshire, England) from its trace metal and organic carbon content. Comparison of the modern
Kimmeridge Clay and Jurassic sediment compositions reveals that the rate of photic zone primary productivity in the
Kimmeridge Sea, based on the accumulation rate of the marine fraction of Ni, was as high as 840 g organic
carbon m− 2 yr−1. This high level was possibly tied to the maximum rise of sea level during the Late Jurassic
that flooded this and other continents sufficiently to allow major open-ocean boundary currents to penetrate
into epeiric seas. Sites of intense upwelling of nutrient-enriched seawater would have been transferred from
the continental margins, their present location, onto the continents. This global flooding event was likely
responsible for deposition of organic matter-enriched sediments in other marine basins of this age, several of
which today host major petroleum source rocks.
Bottom-water redox conditions in the Kimmeridge Sea, deduced from the V:Mo ratio in the marine
fraction of the Kimmeridge Clay Formation, varied from oxic to anoxic, but were predominantly suboxic, or
denitrifying. A high settling flux of organic matter, a result of the high primary productivity, supported a high
rate of bacterial respiration that led to the depletion of O2 in the bottom water. A high rate of burial of labile
organic matter, albeit a low percentage of primary productivity, in turn promoted anoxic conditions in the
sediment pore waters that enhanced retention of trace metals deposited from the water column.
Published by Elsevier B.V.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2. Setting the stage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
2.1. The problem of temporal scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
2.2. Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
2.3. Water-column versus sediment redox. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2.4. Nutrient sources. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
2.4.1. Fluvial nutrient source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
2.4.2. Groundwater nutrient source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
2.4.3. Atmospheric nutrient source. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
2.4.4. Oceanic nutrient sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
2.5. Trace-element sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.5.1. Lithogenous trace-element source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.5.2. Seawater trace-element sources, trace nutrients and redox-sensitive trace metals . . . . . . . . . . . . . . . . . . . . . . . 72
2.5.3. Trace-element accumulation rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
2.5.4. Limitations of modeling trace-element ratios and accumulation rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

⁎ Corresponding author. Tel.: +1 650 329 5187; fax: +1 650 329 5491.
E-mail address: dzpiper@usgs.gov (D.Z. Piper).

0012-8252/$ – see front matter. Published by Elsevier B.V.


doi:10.1016/j.earscirev.2009.03.001
64 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

3. Modern marine environments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76


3.1. The Cariaco Basin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.1.1. Oxidant balance in and residence time of bottom water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.1.2. Trace-metal budgets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.2. The Black Sea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.2.1. Primary productivity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.2.2. Oxidant balance in bottom waters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.2.3. Bottom-water residence time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.2.4. Trace-metal budgets in bottom waters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.3. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4. An ancient marine environment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.1. General description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.2. Lithogenous- and marine-sediment fractions: A trace-element partitioning model . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.3. Biogenous sediment fraction: Primary productivity and rate of upwelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.4. Hydrogenous sediment fraction: water-column redox and residence time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.5. Hydrography of primary cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
5. Summary and conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

1. Introduction the first part, we review nutrient and trace-element sources and
cycles in modern marine basins. Several paleoceanographic scenar-
Black shales are major source rocks for oil deposits, commonly ios have been constructed that make it difficult to advance falsifiable
contain economically viable phosphate deposits, and can host a hypotheses for black shale formation and, in many cases, lead to
variety of trace elements that are potentially highly toxic, making a untenable conclusions. These include (1) the supply of nutrients to
study of these deposits both economically and environmentally marine environments by rivers to account for the enrichment of
important. As indicated in several recent reviews of their environ- organic matter (i.e., high primary productivity) in sediments that
ments of deposition (Tyson and Pearson, 1991; Arthur and Sageman, accumulated under bottom-water O2 depletion, where sub-surface
1994; Wignall, 1994; Harris, 2005), both the organic enrichment and waters and their nutrient inventory were supposedly isolated from
generally high contents of trace and minor elements that form the photic zone; (2) the isolation of anoxic bottom waters to explain
insoluble sulfide phases, or that are reduced under sulfidic condi- diverse occurrences of organic enrichment in sedimentary
tions (Vine and Tourtelot, 1970; Gustafson and Williams, 1981; Jacobs sequences (Woolnough, 1937; Brockamp, 1944; Demaison and
et al., 1985, 1987; Crusius et al., 1996), have been ascribed to accu- Moore, 1980) even though in many deposits sedimentological and
mulation in basins that had anoxic bottom waters. This condition paleontological evidence points to the presence of oxygenated to
is often conjectured to have been maintained by a stratified water denitrifying or, at least ventilated conditions at the seafloor; and
column that isolated the photic zone from the bottom-water nutrient (3) the assumption that high concentrations of organic matter in
reservoir (Hatch and Leventhal, 1992; Turner, 1992; Ettensohn, 1995; these deposits equate to a high degree of its preservation relative to
Wignall and Myers, 1988). Contrary to this interpretation, the organic primary productivity.
richness of many black shales has led some geologists to conclude In the second part, we present a quantitative analysis of organic
that they were deposited under conditions of high primary pro- matter and trace-element fluxes to the seafloor in two modern anoxic
ductivity, together with a relatively low degree of dilution of the basins—the Cariaco Basin and the Black Sea—that provides insight
organic matter by siliciclastic and skeletal, largely planktonic, debris into the relationships between the hydrography, surface-water
(e.g., Lallier-Vergès et al., 1995; Herbin et al., 1995; Perkins et al., primary productivity, and bottom-water redox conditions of the
2008). two basins, and the eventual accumulation of trace elements and
Scenarios for the control of productivity and for the maintenance organic matter on the seafloor. The results establish sedimentary
of anoxic depositional conditions are highly diverse. Most rely on, or trace-element proxies for these three important properties of the
make assumptions about, processes and conditions in the modern water column.
ocean. Thus, a firm understanding of basic oceanographic processes is In the final section, we apply this analysis to an examination of
required if explanations of black shale formation are to be based on oceanographic conditions that prevailed during deposition of the
uniformitarian principles. Such understanding should constrain extensively studied Late Jurassic Kimmeridge Clay Formation (KCF) of
conditions of black shale deposition and advance our overall knowl- western Europe (Cox and Gallois, 1981; Wignall, 1990; Herbin et al.,
edge of the factors that led to the formation of this important lithology 1995; Tyson, 1996; Tribovillard et al., 2005), focusing mainly on the
that occurs repeatedly through geological time. While recognizing geochemistry of samples from three cores from the Cleveland Basin,
that reworking and redeposition of sediment, chemical alteration Yorkshire, England (Fig. 1), but with reference to the geochemistry of
during and following burial, and surficial weathering following uplift samples from outcrops of age correlative formations of the Boulonnais
and exposure all may mask and, in some cases, erase the original Basin in northern France.
geochemical signal of deposition, application of modern oceano- We adopt a uniformitarian approach that focuses on oceanographic
graphic principles and the geochemistry of modern marine deposits processes rather than attempting to advance unique paleogeographic
demonstrate that several currently held interpretations of black shale models for ancient sedimentary basins. Although we examine the
accumulation need to be re-examined. biogeochemical processes in the shallow-silled and largely isolated
In this review, we adopt a three-fold approach to examine Cariaco Basin and Black Sea, we recognize that they may not be
explanations of black shale formation that invoke former levels of appropriate bathymetric models for basins in which black shales
phytoplankton production and bottom-water oxygen depletion. In accumulated in the geologic record. Indeed, our analysis suggests
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 65

Fig. 1. Paleogeographic map of the Late Jurassic Kimmeridge Sea, Western Europe, showing the locations of the two sites sampled for chemical analysis. Chemical analyses were made
available by N. Tribovillard. The map is adapted from Herbin et al. (1995). For a presentation of the bathymetry during the Late Jurassic see Ziegler (1990).

just that. However, understanding the hydrographic, chemical, and on the deposition of the Kimmerdigian organic- and metal-enriched
biological processes of the water column in these modern envi- shales that accumulated from vast, perhaps non-silled and relatively
ronments helps to unravel the geochemical and hydrographic controls shallow, epeiric seas.
66 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

2. Setting the stage that might have extended over several hundred to even thousands of
years, owing to constraints of sampling and sediment accumulation
2.1. The problem of temporal scales rates, whereas modern sediments can be sampled on a scale of less
than a year to decades. For example, in cores recovered from the
Equal thicknesses of modern sediments and ancient rock strata Cleveland Basin (Herbin et al., 1995; Tribovillard et al., 2005), the part
represent different time intervals, owing to effects of sediment com- of the section of the KCF that extends from the top of the Mutabilis
paction as well as the possible incompleteness of the rock record. The ammonite zone to the top of the Wheatleyensis zone is as much as
geochemistry of black shale samples commonly represents conditions 130 m thick (Fig. 2) and represents a time interval of 2.8 to 4 My

Fig. 2. Stratigraphic section of the Kimmerdige Clay Formation in Yorkshire, England (the Cleveland Basin) and of correlative formations at Cap de la Crèche, northern France (the
Boulonnais Basin), adapted from Herbin et al. (1995). The five organic-carbon-rich intervals (ORIs) in the Yorkshire area are based on the distribution of Corg. Their positions at the
Cap de la Crèche site are based largely on paleontologic correlations with the section in Yorkshire, rather than on the actual distribution of Corg (Herbin et al., 1995). Uncertainty of the
positions of the ORIs at this site is depicted by the broken symbols. For the purposes of this paper, the intervening intervals are referred to in the text as organic-poor intervals (OPIs),
each occurring immediately below the corresponding numbered ORI.
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 67

(Gradstein et al., 1995, 2004). We have adopted an interval of 3.6 My fraction incorporated in aluminosilicate lattices) as oxides under this
(Weedon et al. (2004), which yields the following sedimentation rate: redox condition include Fe, Mn (Fig. 3), and several trace elements
adsorbed by the oxyhydroxides of these two metals (Nameroff et al.,
Sediment accumulation rate = 13; 000 cm  3:6 my 2002). However, Fe and Mn are the critically diagnostic elements.
−1
= 0:0036 cm yr : ð1Þ Suboxic conditions are defined by the depletion of oxygen below ca.
50 μM, but prior to the onset of SO2−
4 reduction (Richards, 1975; Glazer
Each 1 cm thick sample represents 280 years of deposition, et al., 2006), when dissolved NO− 3 and the remaining O2 act as the
severely averaging the geochemical conditions of deposition that may major terminal electron acceptors. Solid oxyhydroxides of Mn, Fe, and
have changed on time scales of decades or even seasons. At an average Ce and dissolved species of Cr, Re, U, and V act as trace, subordinate
density of 2.3 g cm− 3 (Weedon et al., 2004), the mass accumulation electron acceptors (Fig. 3). The latter four metals are reduced to less
rate was 83 g m− 2 yr− 1. soluble ionic states than under oxic conditions and eventually
accumulate on the seafloor. Because of the much lower concentration
2.2. Definitions of NO− 3 in seawater and, hence, the narrow redox range over which it
is the dominant electron acceptor, suboxia may not have been
We adopt the following definitions and terminology to categorize considered significant in the geological record. However, it is the
marine environments having different dissolved oxygen concentra- dominant redox condition under which: (1) the essential and, in many
tions. The boundaries of oxygenation regimes (Fig. 3) are set by the areas of the ocean, limiting nutrient NO− 3 is lost from the dissolved
sequential microbially-driven oxidation reactions that degrade fixed-N pool (N2O, NH3, NO− −
2 , and NO3 ) in seawater (Codispoti, 1989),
organic materials in marine waters and sediments (Breck, 1974). (2) carbonate-fluorapatite accumulates (Burnett and Froelich, 1988),
Oxygen is the major oxidant in all surface marine waters that are in and (3) several major sedimentary phosphorite deposits (Fonseca,
equilibrium with the atmosphere, and in sub-surface waters and 2000; Piper, 2001) and black shales may have accumulated in the past
sediment pore waters where oxygen is above a concentration of ca. (Perkins et al., 2008). Anoxic conditions ensue when the reservoirs of
50 μM (Berner, 1981). Within the range of 50–300 µM (depending on these other oxidants have been essentially exhausted and H2S is
temperature) organic detritus will be oxidized by oxygen as the produced by SO2− 4 reducing bacteria (H2S is used throughout the
terminal electron acceptor under oxic conditions. Trace elements that paper to refer to all forms of dissolved sulfide). Apparent reduction of
accumulate from seawater (i.e., not including the generally dominant MoO2− 4 and Cu2+ and their accumulation on the seafloor, along with

Fig. 3. Schematic representation of bacterial respiration versus water-column depth in marine basins experiencing O2 depletion in the bottom water. Profiles of O2, NO− −
3 , NO2 , NH3
(NH3 plus NH+ 4 ) and H2S (all forms of dissolved S
2−
) concentrations are representative of all anoxic basins, with depths, zonal thicknesses, maximum oxidant concentrations varying
from basin to basin. Maximum values for each in the Cariaco Basin are ca. O2 210 μM, NO− −
3 12 μM, NO2 0.1 μM, H2S 40 μM, and NH3 15 μM (Richards, 1975). Half-cell redox reactions are
shown for the major electron acceptors. The different colors are intended to identify the oxic photic (light blue) and aphotic zones (dark blue), suboxic zone (orange), and anoxic
zone (green). Approximate thicknesses of redox zones in various modern basins depleted in O2 at depth are as follows: oxic (50–150 m), suboxic (50–150 m), and anoxic (200 to
2100 m). Stabilities of trace-element species under the different redox conditions are shown in the center, calculated from thermodynamic constants (see Piper, 2001). In the case of
Cr, it precipitates possibly as Cr(OH)3, or is adsorbed onto settling particles, under mildly denitrifying to anoxic conditions, i.e., throughout the O2 depleted region of the water
column. Under oxic conditions, Cr is in the oxidized and more soluble CrO2− 4 valence state (Murray et al., 1983). MnO2 responds oppositely; it is reduced to a soluble valence state
(probably Mn2+) under anoxic conditions, but is stable within the oxic realm. Depths of several modern basins are shown at far right, together with the estimated depth of the
Jurassic Cleveland Basin. Redox conditions in the water columns of each are given by the vertical lines, for which the broken lines represent temporally varying redox conditions in the
bottom waters.
68 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

the reduced species of Cr, Re, U, and V and precipitates of FeS, CdS and The establishment of redox conditions in modern marine sedi-
ZnS (Fig. 3) characterize this level of bacterial respiration (Jacobs et al., ments depends on the settling flux of particulate organic matter to the
1985, 1987). seafloor and its rate of burial (Canfield, 1989, 1994). The settling flux
Other terms have been used to describe oxygenation regimes through the water column decreases with water depth (Suess, 1980;
in marine environments (Tyson, 1987). Dysoxic conditions have Pace et al., 1987; Thunell et al., 2000; Goñi et al., 2003), such that a
been defined for environments where dissolved oxygen levels are high burial flux of organic carbon (Corg), relative to primary
ca. 10–100 µM, encompassing much of the suboxic range defined by productivity, is only possible at seafloor depths of roughly hundreds
the geochemistry of the deposit. Dysoxic conditions equate to the of meters versus thousands of meters that characterize the pelagic
dysaerobic biofacies. Hypoxic conditions are normally used to describe environment. Within modern sediment, the boundary between oxic
variably reduced concentrations of dissolved oxygen that cause stress and anoxic zones varies from less than a millimeter to greater than
in aquatic organisms. These finer-scale distinctions have been erected 15 cm below the seafloor when progressing from coastal embayments,
on the basis of distinctive benthic faunal assemblages rather than the to the continental shelf, to the upper slope, and finally to the lower
presence/absence of specific electron acceptors. The use of the suffix continental slope.
aerobic is confined to the physiology of the organism or consortium The oxic–anoxic boundary, based on Eh measurements and trace-
that utilizes a particular oxidant. Exaerobic is also used as a biofacies element inventories (Bonatti et al., 1971), is further defined by a color
term for environments where shelly macrofauna are found on anoxic change from dark brown at the surface to light green at depth (Lyle,
sediment surfaces that are not bioturbated, and where oxygen levels 1983). In many shallow sea-floor environments, intensely anoxic
in the bottom water and surface sediment pore water are ca. 5–10 µM conditions (high dissolved sulfide in the pore waters) are commonly
(Savrda and Bottjer, 1987). Thus, aerobic, dysaerobic and anaerobic observed at or within a few millimeters of the sediment–water inter-
facies equate to oxic, suboxic, and anoxic regimes (Tyson and Pearson, face, above which fully oxic conditions prevail. Bonatti et al. (1971)
1991). The latter are more useful for geochemical studies of sediments reasoned that increasing rates of bulk-sediment accumulation from
as they are defined uniquely by the suite of trace elements that lower slope to shallow shelf allowed for increasing burial rates of
accumulates on the seafloor under the specific redox conditions of the labile organic matter in nearshore settings. A decreasing flux of
bottom water. The term euxinic (Black Sea-like, from the Greek Pontos organic matter to the seafloor, owing to a decrease in primary
Euxinus; King, 2004), commonly used in the geological literature, is productivity away from the coast (Mullin, 1986) and increasing ocean
used to depict sulfidic bottom waters as distinct from those situations depth (Suess, 1980), also contribute to this change. Organic matter is
where sedimentary pore waters are anoxic (sulfidic) but measurable further degraded during burial (Dean and Gardner, 1998), but largely
oxygen is found in the bottom waters (Berner, 1981). Tyson (1995) in the uppermost centimeters under the redox conditions of the
suggests that anoxic be used to describe sulfidic bottom-water bottom waters (Jørgensen, 1982). Where sediment accumulation
conditions, which we adopt. rates are high, a relatively large fraction of labile organic matter
escapes oxidation at the very surface (Canfield, 1989, 1994; Betts and
2.3. Water-column versus sediment redox Holland, 1991). Its oxidation at depth in the sediment depletes the
pore waters of residual O2, NO− 3 , and Mn/Fe oxyhydroxides, thereby
The attribution of preferential organic matter accumulation in promoting SO2− 4 reduction near and even at the sediment surface.
ancient rock sequences to anoxia frequently fails to distinguish Under this condition, organic matter accumulates (is buried) under
explicitly between paleoceanographic settings where the bottom anoxic conditions, but was deposited under oxic water-column con-
water of a basin is anoxic, and therefore the surface sediments ditions. Although the sediments in this situation accumulate beneath
themselves are anoxic, and where bottom water immediately above an oxic water column, they may contain mineral phases, e.g., pyrite, or
the sediment/water interface is suboxic or even oxic, but the sediment elemental species that can only form or be fixed under low redox
pore water is anoxic (Tourtelot, 1979). Examples of the first scenario potentials. They are therefore archives of accumulation under anoxic
are found in several well-known basins characterized by a physical conditions. The distinction between water-column and sediment
barrier, which restricts bottom-water renewal or ventilation, such that pore-water anoxia is critically relevant to any discussion of the envi-
the oxygen content of the bottom water becomes depleted when the ronment of formation of black shales, in view of the emphasis com-
supply rate of reduced (organic) carbon exceeds the flux of oxygen monly placed on anoxic water-column environments as a necessary
from outside the basin. Bacterial respiration within the basin water condition for the formation of this facies.
(Fig. 3) is driven first by aerobic metabolism of settling organic debris Maintenance of the different conditions of bacterial respiration in
using oxygen as a terminal electron acceptor, followed by NO− 3 the water column depends on the balance between the supply of the
reduction, and finally SO2− 4 reduction. This depth sequence is reductant Corg and the oxidants O2, NO− 2−
3 , and SO4 . As described
observed in the water column of all basins in which anoxic conditions previously, O2 is utilized by bacteria in the remineralization of Corg.
are present in the bottom water (Glazer et al., 2006). Oxygen is supplied by mixing of near-surface and deeper waters and
The oxygen minimum zone (OMZ) in some areas of the modern by the sinking and lateral flow of oxygenated waters below the near-
ocean has dissolved O2 contents that are a few percent of surface- surface waters. Vertical mixing is often severely inhibited by the
water concentrations (Karstensen et al., 2008), which remain vertical density stratification of ocean waters, in turn controlled by
oxygenated due to near-surface photosynthesis and gas exchange vertical gradients of temperature and salinity. The lateral circulation,
with the atmosphere. The low oxygen conditions at depth are con- or advection, of water along constant density (isopycnal) surfaces is a
trolled by the rate of lateral intermediate water circulation (Wyrtki, more important mode of intermediate and bottom-water ventilation.
1962), balanced against the rate of bacterial respiration, which is An illuminating example of advection is found on the continental
determined by the settling flux of organic matter from local surface margin of Peru, where the poleward Peru Undercurrent injects
waters and much less so from local river sources (see Section 2.4.1). oxygenated seawater across the seafloor within the OMZ where it
Extensive areas in the eastern tropical Pacific, the Arabian Sea, and the impinges on the outer shelf and upper slope. Although bottom-water
southeastern Atlantic have dissolved O2 concentrations b5 μM; these oxygen concentrations are a few micromolar, the vigorous flow of
waters are suboxic (Codispoti, 1980; Murray et al., 1983; Codispoti this water, albeit with very low O2 concentrations, maintains a robust
et al., 2001; Nameroff et al., 2002; Kuypers et al., 2005; Morford et al., macrofaunal community (Arthur et al., 1998). Likewise, oxygen is
2005). Hydrogen sulfide has been reported in some OMZ (Ivanenkov injected into shallow intermediate anoxic waters in the Black Sea as
and Rozanov, 1961; Dugdale et al., 1977), but anoxic conditions in the manifested by the interleaving of water with hydrographic properties
open ocean are rare. distinct from the ambient water (Murray et al., 1991; Özsoy et al.,
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 69

1991; Grégoire and Stanev, 2001; Özsoy et al., 2001; Murray et al., the Black Sea have higher Corg:Spyrite ratios than “normal” marine
2007). sediments, with no intercept on the S-axis. Moreover, data from the
The occurrence of pyrite in black shales has often been used as anoxic basins of the Baltic Sea (Boesen and Postma, 1988) and the
evidence for their accumulation under anoxic bottom-water condi- Cariaco Basin (Lyons et al., 2003) also show no systematic S enrich-
tions (e.g. Berry et al., 1989; Buggisch, 1991; Suzuki et al., 1998; ment relative to Corg compared with “normal” marine sediments. In
Caplan and Bustin, 2001; Galeotti et al., 2003; Martinez, 2003). While addition, the sediments accumulating on the continental margins of
sulfidic conditions are required for pyrite formation (Berner, 1970), Peru, Namibia and Oman, which are bathed in low oxygen waters of
the process can occur in sediments that accumulate under fully the OMZ, also have much higher Corg:Spyrite ratios than “normal”
oxygenated water columns as well as in anoxic basins (Goldhaber and marine sediments (Morse and Emeis, 1990; Suits and Arthur, 2000).
Kaplan, 1974; Goldhaber et al., 1977). Thus, fine-grained shallow- The lack of S enrichment under these conditions is due to Fe limitation
water marine sediments characteristically have relatively high during pyrite formation (Lyons and Berner, 1992), owing to the limited
pyrite contents, especially where settling fluxes of organic matter burial of easily reducible solid-phase Fe in siliciclastic minerals and
are high due to high productivity and shallow depths (Suess, 1980). oxyhydroxides, in the latter case because of their reduction to soluble
The concentration of this authigenic phase commonly correlates Fe2+ in the overlying suboxic and anoxic waters. The proposition that
strongly with Corg content due to organic matter limitation of post- sediments accumulating in basins with sulfidic water columns have
depositional sulfide production. The occurrence of pyrite in black distinctively different Corg:S ratios from those sediments that are
shales should not, therefore, be used as an indicator of a sulfidic oxygenated at the sediment/water interface but anoxic at depth is not
(anoxic) basin. supported by these more recent observations.
The Corg:Spyrite ratio has also been used to infer an anoxic water
column in the geological record. This proxy, developed by Leventhal 2.4. Nutrient sources
(1983) using data from recent Black Sea sediments published by Hirst
(1974), noted an intercept on the S-axis in the regression of total S Most primary production in the ocean is confined to near-surface
versus Corg, in contrast to data from “normal” marine sediments sensu waters, where sufficient light is available for photosynthesis. A minor
Berner (1982) that are anoxic but accumulate beneath oxic waters. site of production on a global scale is found in the deep ocean around
Pyrite, i.e. S, enrichment was attributed to the presence of H2S in the hydrothermal vents, where the H2S/SO2− 4 system provides an energy
water column of the Black Sea as well as the sediment pore waters. source (van Dover, 2000). The utilization of essential nutrients
The systematics of this additional or syngenetic pyrite versus the (various chemical species of N, P, Si, and several trace nutrients)
diagenetic pyrite that forms in the sediment has been examined by during phytoplankton growth depletes their concentrations to low
Raiswell and Berner (1985). On the basis of these results, low Corg:S levels over most of the ocean surface (Fig. 4). Regeneration of the
ratios have become a common criterion to identify anoxic water- nutrients occurs in sub-surface waters when particulate organic
column conditions in the geological record (e.g. Leckie et al., 1990; matter settles into deeper waters, resulting in the nutrients extracted
Jones and Manning, 1994). Subsequent work on Black Sea sediments from the near-surface waters increasing in concentration below the
(Calvert and Karlin, 1991; Lyons and Berner, 1992), however, has failed photic zone. They are only very weakly utilized by bacteria in the
to replicate the high S values reported by Hirst (1974). The more deeper water. Sustained primary production is only possible when
recent work shows that the data for Late Holocene sediments of and where nutrients are re-supplied to the surface layers, either from

Fig. 4. Depth profiles of O2, major nutrients, and trace elements in the central North Pacific Ocean. The O2 minimum in the far eastern Pacific is an order of magnitude shallower and an
order of magnitude more intense than shown here and is one of several sites in the open ocean that experiences denitrification in the water column. Except for Mo, the trace elements
exhibit a nutrient-type profile. The uptake of V by algae is discernable, albeit weakly, by its slight depletion in the photic zone (Collier, 1984). For Mo, its high concentration in
seawater (Collier, 1985) and weak uptake by algae (Brumsack, 1986) mask any possible biological depletion in the photic zone and concomitant enrichment at depth via release from
organic matter through the oxidation of settling organic matter.
70 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

land runoff, groundwater flow, atmospheric deposition, or from the in the plume regions and supplied from upwelling of water below the
vertical movement and mixing of sub-surface waters into the surface surface fresh-water plumes (Corredor et al., 2003).
photic layer of the ocean. Nutrient delivery to the Atlantic by the Zaire River also appears to
have only a modest impact on production in the coastal region off
2.4.1. Fluvial nutrient source western Africa. Production is lowest within the offshore region where
The limitation of marine production by the supply of nutrients to salinities show a 50/50 mixture of river water and seawater (Cadée,
the photic zone, due to strong water-column stratification as is found 1978). It is highest at zero salinity and at normal salinity, i.e.,
in anoxic basins, has commonly led to the attribution of a dominantly undiluted, offshore seawaters. The higher offshore production is due
fluvial nutrient input to a sedimentary basin to explain the high once again to coastal upwelling, which is highest in summer (van
organic content of some black shales (Baird and Brett, 1991; Bennekom et al., 1978) when the eastern boundary surface Benguela
Ettensohn, 1995). Terrestrial rocks and soils are the ultimate sources Current is strongest, with the Coriolis component of flow (Ekman
of the nutrients phosphate and silica (HPO2− 4 and H4SiO4 at sea water transport) offshore to the west, or to the left of the direction of the
pH), which are delivered to the ocean in dissolved forms via rivers. dominant flow (Southern Hemisphere).
Although some N is also supplied to the ocean via rivers, the principal Nutrient levels in estuaries are often higher in waters at intermediate
source of N to the ocean is from the fixation of N2 by diazotrophic salinities than in either the river or the offshore end members. This well-
cyanobacteria within the ocean (Karl et al., 2002). Nitrate (NO− 3 ), known phenomenon reflects the two-layer estuarine flow of the waters
the main form of fixed-N in the open ocean, is produced by and the recycling of nutrients between particulate and dissolved forms.
the remineralization of organic-N via the oxidation of NH3 (nitrifica- Organic debris that falls into deeper waters can be retained within the
tion) liberated from proteins. Thus, the cycles of N and P, the two estuary as a consequence of the compensating return flow of water from
macronutrients essential for phytoplankton growth, are de-coupled. offshore below the out-flowing fresher surface waters. Remineralization
Dissolved Fe also limits primary production in some oceanic of this organic matter releases nutrients to the dissolved inventory,
regions that have unutilized nutrients in the photic zone. It is a co- leading to their accumulation in the inflowing seawater below the
factor in the enzyme NO− 3 reductase, required by phytoplankton cells surface flow. Such recycling causes estuaries to be nutrient traps
to reduce NO− 3 to ammonium before incorporation in protein syn- (Redfield et al., 1963), or filters. Entrainment of deeper waters into the
thesis. Large areas of the Southern Ocean, the Subarctic and equatorial out-flowing surface waters at the boundary between the two opposing
Pacific, the so-called High Nutrient Low Chlorophyll (HNLC) regimes flows enriches the surface sun-lit waters with the regenerated nutrients
of the open ocean, have inadequate Fe supplies either from deeper and with nutrients transported into estuaries from offshore, resulting in
waters, coastal lands, rivers, or bottom sediments, relative to the estuaries often being much more productive than open-ocean regimes.
supplies of NO− 2−
3 and HPO4 . Efficient utilization of the ambient For example, surface water fixed-N, together with organic-N and HPO2− 4
nutrients (Martin, 1991, 1992) is limited, owing to Fe precipitation as concentrations in the Amazon River estuary are highest at salinities
an oxyhydroxide in the presence of oxygen, which effectively removes between 2 and 8‰ compared with the river end-member and offshore
it from seawater. Several coastal regions also have HNLC character- waters at oceanic salinity (Edmond et al., 1981). Other major estuaries,
istics, as off northern California (Hutchins and Burland, 1998) and such as the Zaire (van Bennekom et al., 1978), the St. Lawrence (Coote
Peru (Hutchins et al., 2002). and Yeats, 1979), and the Changjiang (Edmond et al., 1985), are similarly
Available information on the impact of fluvial runoff on ocean enriched in nutrients at intermediate salinities. In some estuaries, the
production suggests that the stimulation of production by this source process of entrainment can supply significantly more nutrients from
is confined to estuaries or to fairly restricted areas of the coastal ocean. offshore than that delivered by the river itself (Yin et al.,1997). Estimates
In the case of the Amazon River, which is responsible for approxi- of the supply of nutrients to the continental margin of the North Atlantic
mately 18% of the global fresh-water runoff to the ocean (Gibbs, 1967) show that large rivers and estuaries currently supply only around 10–
and is affected only to a minor degree by anthropogenic nutrient 15% of the total (dissolved and particulate) fixed-N and HPO2− 4 to the
additions, most of the dissolved fixed-N and HPO2− 4 carried by the shelf region, whereas the offshore supplies 80–90% (Nixon et al., 1996).
river is removed within and immediately off the river mouth when The larger offshore oceanic nutrient supply is due to the estuarine
most of the suspended load has been deposited and sufficient light circulation of most shelf systems; low-salinity surface water flows
penetration into the waters promotes estuarine algal production offshore to be replaced by nutrient-rich deep water that is sourced
(Edmond et al., 1981; DeMaster et al., 1996; Smith and DeMaster, beyond the shelf break (Iselin, 1939; Hseuh and O'Brien, 1971;
1996). Between 25 and 60% of the river-borne dissolved silicon is Wroblewski and Hofmann, 1989). This process, as noted earlier, even
removed by the growth of diatoms also within the river plume leads to the import of deep-water nutrients into the surface plumes of
(Milliman and Boyle, 1975; DeMaster and Pope, 1996; DeMaster et al., rivers (Geyer et al., 1996). In the Amazon, that supply is 5- to 10-fold
1996). Although high production rates characterize the shelf waters greater than the supply from the river itself (DeMaster and Pope, 1996).
close to the Amazon River mouth due to a combination of fluvial and Despite the high productivity of shelf waters off the Amazon River
offshore nutrient supplies (see Section 2.4.4), the low-salinity waters that is only partly due to fluvial nutrient supply, the sediments of the
flowing northwards along the coasts of Brazil and the Guyanas are shelf affected by the river plume are composed mainly of terrigenous
severely depleted in nutrients (Ryther et al., 1967; DeMaster et al., silty clays carried to the shelf predominantly in suspension (Kuehl
1996). Coastal production is seasonally higher than it is offshore, but it et al., 1996; Sternberg et al., 1996). The organic matter and biogenous
results from upwelling (see Section 2.4.4) of nutrient-enriched silica produced in the river plume and deposited on the shelf are
marine waters between the coast and the Amazon plume. This is rapidly remineralized in the strongly reworked bottom sediments
due to the divergence of surface waters (Ekman transport) away from (Aller et al., 1996) and diluted by the high lithogenous background;
the coastal barrier during summer (Gibbs, 1980; Geyer et al., 1996), Corg contents of the sediment are b1% by weight (Aller and Blair,
when the Intertropical Convergence Zone (ITCZ) lies north of the 1996), with as much as 70–90% of terrestrial origin (Showers and
equator (Haug et al., 2001), causing the southeast trade winds to blow Angle, 1986; Sommerfield et al., 1996). Thus, as is the case in large
along the coastline. Although fluvial sources of surface waters river deltas and their offshore extensions, most of the organic matter
probably derived from the Amazon and Orinoco Rivers can be is derived from the land rather than having a marine source (Berner,
detected in the western tropical Atlantic (Müller-Karger et al., 1982; Hedges and Keil, 1995).
1989b) and the Caribbean (Moore et al., 1986; Müller-Karger et al., We have avoided a discussion of an anthropogenic input of
1989a; Bonilla et al., 1993), production in these areas is largely nutrients to rivers, although this is an important source currently that
confined to frontal regions and is based entirely on nutrients recycled can lead to major environmental problems, particularly in coastal
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 71

areas (e.g., van Bennekom et al., 1975; Maybeck, 1982; Halim, 1991; Upwelling of sub-surface water also takes place in the Antarctic
Rabalais et al., 1996; Lancelot et al., 1997). Circumpolar Current with the transport of surface waters away from
the continental boundary due to the westerly winds (Gordon, 1971).
2.4.2. Groundwater nutrient source Upwelled water along coastlines and the equator is drawn to the
Groundwater seepage also supplies fixed-N and HPO2− 4 to estuarine surface from depths not exceeding 200 to 300 m (Sverdrup et al.,
waters and directly to coastal regions (Johannes, 1980; Capone and 1942) and, in some cases, shallower than 100 m depth (Codispoti,
Bautista, 1985; Giblin and Gaines, 1990; Varela et al., 1990). This source 1980). In spite of the shallow source, this upwelled water has a higher
has become more important in recent decades owing to contamination nutrient content than surface water, although not as high as abyssal
of groundwaters by municipal and agricultural runoff (Nolan and Stoner, waters (Fig. 4). The moderately higher contents result from the first
2000). However, groundwaters are also enriched in nutrients even in stages of the remineralization of labile organic matter settling to depth
undisturbed forested watersheds (Ford and Naiman, 1989) so that this from the surface productive layer of the sea and by lateral transport
source was probably important prior to the rise of human populations. In from the Southern Ocean. The vertical flow, or upwelling, of this water
some coastal areas, the flux of nutrients via groundwater seepage is of then serves to replenish the surface layers with regenerated nutrients
the same order of magnitude as that delivered by fluvial runoff (Chavez and Smith, 1995; Hutchings et al., 1995).
(Rutkowski et al., 1999), the higher concentrations of nutrients Nutrient inventories in shallow sub-surface waters of the ocean
compensating for the lower flow rates compared with riverine flows. vary on climatic timescales that are controlled by the mode of
Although groundwater inputs may contribute significantly to primary circulation of the ocean's deep overturning circulation (Toggweiler
productivity in some local areas (Burnett et al., 2003; Slomp and Van et al., 2006). Around Antarctica, the eastward flowing surface
Cappellen, 2004), reliable estimates of this nutrient source to the ocean circulation diverges northwards away from the continent, inducing
on a global scale are very limited because groundwater flow into the large-scale upwelling of deep water. The upwelled water is relatively
ocean is diffuse and patchy. nutrient-enriched because it is derived from the southern branch of
the deep overturning circulation where regenerated nutrients from
2.4.3. Atmospheric nutrient source the entire Atlantic have accumulated during the southward flow of
An atmospheric supply of fixed-N to estuaries and the coastal deep water. The upwelled water now at the surface is driven
ocean has increased markedly over the last several decades, again as a northwards by Ekman transport. After surface cooling and vertical
consequence of industrial and agricultural activities (Paerl, 1993, convective homogenization during the winter, it sinks to shallow
1995). Although the impact of this source is currently important in intermediate depths as the Subantarctic Mode Water (McCartney,
areas downwind of major population centers, it was not important 1977) and circulates through the thermoclines of all ocean basins as
compared with natural sources prior to the anthropogenic era. Antarctic Intermediate Water. This supply of sub-thermocline waters
to all ocean basins delivers nutrients that ultimately drive low-latitude
2.4.4. Oceanic nutrient sources primary production (Sarmiento et al., 2004).
The impacts of fluvial, groundwater, and atmospheric sources of Nutrient supply by this mechanism probably changed on glacial–
nutrients to the ocean pale in comparison to the supply of nutrients to interglacial timescales due to changes in ocean stratification and/or
surface waters from the ocean interior. Waters deeper than the surface changes in Southern Ocean upwelling. Glacial deep ocean waters were
mixed photic layer of the ocean have higher concentrations of nu- denser than modern abyssal waters because of intense sea ice
trients than surface waters. More than 90% of the nutrients extracted formation near Antarctica. This slowed vertical mixing and conse-
from the photic zone by an opportunistic biological community are quently reduced upwelling (Watson and Naveira Garabato, 2006).
returned to solution at depth by microbial remineralization (Fig. 4). Upwelling could also have been reduced by the weakening and
They are then returned to the photic zone by advection, vertical northward migration by 7°–10° latitude of the circum-Antarctic wind
mixing, and diffusion. This biological pump (McElroy, 1983; Knox and belt (Toggweiler et al., 2006). This would have suppressed upwelling
McElroy, 1984) maintains the concentration gradient between surface of deep water near Antarctica leading to decreased nutrient re-supply
and deep waters (Sarmiento and Gruber, 2006), opposing the to thermocline waters. As thermocline depths are the source regions
tendency for the uniform distribution of nutrients in ocean waters for coastal and equatorial upwelling, nutrient supply to surface waters
by advection and mixing. and hence productivity would have been modulated by glacial events.
At high latitudes, surface- and deep-water nutrient levels are Plankton nutrition prior to the Pleistocene could also have been
homogenized during winter months when production is light limited. affected on other time scales, e.g., long-period changes in sea level
Warming in spring, which stabilizes surface waters, sets the condi- (Haq et al., 1987; Miller et al., 2005) that resulted in flooding events on
tions for the spring algal bloom when there is sufficient light for continental margins and changes in atmospheric temperatures
photosynthesis (Mann and Lazier, 2006). At lower latitudes, surface (Scotese, 2001).
waters are re-fertilized by mixing across the pycnocline (diapycnal Although the ultimate sources of all nutrients to ocean waters are
mixing) and by the upward flow of waters from below the the atmosphere and continents, the proximate nutrient supply to
thermocline, a process known as upwelling (Sverdrup et al., 1942). surface waters is dominantly from within the ocean itself, namely
Upwelling is prevalent along continental margins that experience relatively shallow sub-surface waters, whence nutrients are entrained
across-shelf transport of surface water and in the open ocean along into the photic zone by mixing and by upwelling. A recent modeling
the equator. study of nutrient supply to surface waters of the North Atlantic
Upwelling is driven by divergence of surface waters (equatorial (Oschlies, 2002) has shown that vertical mixing is the dominant
case; Sverdrup et al., 1942) and by transport of surface waters away supply mechanism in the subpolar region, while upwelling is most
from a coastline (Wyrtki, 1962; Smith, 1995), expressions of Ekman important in equatorial and coastal regions. Upwelling alone, in the
transport. Continuity requires the transport of surface water to be coastal, equatorial, subpolar, and monsoonal systems, accounts for
compensated by the upwelling of sub-surface water, with its higher around three quarters of the new production in the ocean (Eppley and
nutrient content than the surface water it replaces (Fig. 4). Upwelling Peterson, 1979), that production fueled by nutrient supply from
occurs at the equator in the Pacific Ocean due to the change in the outside the surface layer (Chavez and Toggweiler, 1995). These two
Coriolis deflection as the Trade Winds reach across the geographical source vectors, together with the low-intensity diapycnal mixing
equator (the meteorological and geographical equators do not across the strong thermocline in the central subtropical gyres, sustain
coincide), and between juxtaposed surface currents flowing in almost all “new” production in the ocean as a whole, dwarfing local
opposite directions north of the equator (Sverdrup et al., 1942). supplies from the atmosphere, rivers, and ground waters.
72 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

2.5. Trace-element sources We use a slightly different procedure to establish the composition
of this fraction in any sedimentary deposit than taking the
The sources of trace elements to the oceans are (1) the weathering composition of a single standard as the model composition of the
and erosion of the continents, which supply trace elements to rivers lithogenous fraction, although the standards provide a check on the
and the atmosphere (Rex and Goldberg, 1958; Martin and Meybeck, final model composition. Trace-element and major-element-oxide
1979), and (2) submarine hydrothermal activity, largely in the deep ratios used to evaluate the composition of the lithogenous fraction
ocean (Lyle, 1976; Corliss et al., 1979; Von Damm, 1990). For most include K2O:Al2O3, Th:Al2O3, Rb:Al2O3, Co:Al2O3, and Ga:Al2O3. All of
trace elements, their transit time in and mixing rates of the ocean these major and trace elements have a dominantly lithogenous
preclude identification of local sources. Trace elements in marine siliciclastic source, thus the constancy of such ratios in a series of
deposits can therefore be viewed as having either a lithogenous sediments or rock samples provides support for the use of a single
source or a seawater source (Goldberg, 1963). Marine sources include composition for this fraction. The similarity of the ratios to that of a
seawater, whereby trace elements are removed from the water single standard, such as a World Shale Average (WSA) and
column as precipitates and adsorbed phases and/or by diffusion continental crust (CC) (Turekian and Wedepohl, 1961; Wedepohl
across the benthic boundary (hydrogenous fraction), and particulate et al., 1969–1978; Wedepohl, 1971), upper continental crust (UCC)
organic matter that settles onto the seafloor mostly from the photic (Taylor and McLennan, 1985; Wedepohl, 1995), post-Archean
zone, having escaped oxidation in the water column (biogenous Average Shale (PAAS) (Taylor and McLennan, 1985), or North
fraction). American Shale Composite (NASC) (Gromet et al., 1984; Condie,
Advances in our understanding of the thermodynamic properties 1992) further supports the selection of a single composition for this
of seawater, trace metals, and marine mineral phases (Sillén, 1961; sediment fraction (Piper and Isaacs, 1994, 1996; Piper and Dean,
Garrels and Christ, 1965, Stumm and Morgan, 1996; Millero, 2001) 2002; Piper et al., 2007; Perkins et al., 2008) in any single deposit.
made since the middle of the last century have provided the means to The lithogenous contribution of elements that may be enriched
hindcast depositional conditions in the rock record. Insight into the above their concentration in the lithogenous source-rock model is also
factors leading to individual element enrichments and depletions in a determined by normalizing bulk concentrations to Al2O3 (Piper and
given sedimentary deposit was further improved through the rapid Isaacs, 1996; Perkins et al., 2008; Calvert and Pedersen, 2007). On a
development within the past 45 years of new analytical techniques for plot of element versus Al2O3, a curve drawn from the origin through
determining trace element and organic matter concentrations in samples with the lowest concentration of the element in question
seawater and modern marine sediments from a wide spectrum of establishes the element:Al2O3 ratio of the lithogenous model. This
depositional environments (e.g., Brewer and Spencer, 1974; Boyle value, or the slope of the curve, is again compared to element:Al2O3
et al., 1976, 1977; Bacon et al., 1980; Elderfield and Greaves, 1982; ratios of the shale standards. In the case of the KCF, the slopes of these
Jacobs and Emerson, 1982; Bruland, 1983; Kremling, 1983; Jacobs et al., minimum curves for Ni, V, and Mo versus Al2O3 are within 10% of their
1985, 1987; DeBaar et al., 1988; Emerson and Huested, 1991; Landing ratios in the WSA (Fig. 5), well within the variations in composition
and Lewis, 1991; Piper and Wandless, 1992; German et al., 1993; between the different shale standards. Copper and Zn have slopes that
Sholkovitz, 1993; Crusius et al., 1996; Calvert et al., 2001; Hedges et al., are roughly two-thirds their respective ratios in the WSA, but close to
2001). A major result of this work has been the precise analysis of mean ratios in the UC and CC standards (Wedepohl, 1995). The con-
those trace elements whose accumulation in marine sediments and centration of Al2O3 in a sample, together with the element:Al2O3 ratio,
sedimentary rocks is diagnostic (Fig. 3) of deposition under oxic (Mn provides an estimate of the lithogenous contribution of the element
and Fe), suboxic (Cr, Re, U, and V), and fully anoxic bottom-water itself.
conditions (Mo, Cd, Cu, Zn). Similarly, knowledge of the marine The relationship between Hf and Zr, both of which have a solely
biogeochemical cycling of trace nutrients by plankton (Fig. 4), e.g., Ba, lithogenous source, serves several purposes in evaluating the nor-
Cd, Cu, Ni, and Zn (Martin and Knauer, 1973; Collier and Edmond, mative scheme. The strong correlation between these two elements
1983; Brumsack, 1986; Whitfield, 2001; Ho et al., 2003) has (Fig. 6) reflects their close chemical similarity (Hampel, 1968) and,
contributed to an improved insight into the role of primary produc- thus the quality of the analyses and the constancy of composition
tivity in cycling trace nutrients through the ocean and their eventual of the lithogenous fraction in any single deposit. The slightly different
accumulation on the seafloor. Hf:Zr ratios in various “standards”, however, demonstrate the
problem of relying on the composition of a single sedimentary rock
2.5.1. Lithogenous trace-element source standard as the model composition for the lithogenous fraction of all
The contribution of major and trace elements to a deposit by the deposits.
lithogenous source, largely siliciclastic debris, can be modeled from The alternative method of identifying the lithogenous fraction is by
bulk-sediment analyses (Leinen, 1977; Calvert and Price, 1983; Piper separating operationally defined fractions of a sediment sample by
and Isaacs, 1994, 1996), or by chemical extractions (Chester and chemical leaching methods (Chester and Hughes, 1967; Tessier et al.,
Hughes, 1967; Tessier et al., 1979). In the modeling approach, the 1979). One aim is to leave the lithogenous material relatively
lithogenous contribution of an element is determined by normalizing unaffected, thus isolating the soluble fraction(s) that may represent
its concentration to a major-element oxide that is hosted solely in the the marine fraction of a trace element from the lithogenous con-
siliciclastic sediment fraction (Calvert and Pedersen, 2007). Alumina tribution. Incomplete extraction of a given component or the simul-
(Al2O3) is frequently used for this purpose as it constitutes one of the taneous extraction of portions of two or more fractions, depending on
essential building blocks of aluminosilicate mineral phases and its the choice of extracting reagents, can bias these results (Sholkovitz,
concentration is similar in acidic to basic extrusive and intrusive 1989; Nirel and Morel, 1990; Tessier and Campbell, 1991). The success
igneous rocks, most metamorphic rocks, the bulk upper crust, and of this procedure in isolating the marine fraction can be evaluated by
shales (Wedepohl, 1971; Taylor and McLennan, 1985; Wedepohl, an analysis of the insoluble fraction and its similarity in composition,
1995). However, the total composition of the different standards of again, to a shale standard (Piper and Wandless, 1992). Both methods
siliciclastic debris varies considerably, reflecting both real and have their obvious limitations.
procedural differences (Wedepohl, 1995). For example, titanium has
been used in a similar way to (Al2O3), but its content varies more 2.5.2. Seawater trace-element sources, trace nutrients and redox-sensitive
widely in different rock types, and it can be enriched in coarse-grained trace metals
fractions of sediments during deposition as it occurs in Ti-rich heavy In the modeling scheme, trace-element concentrations that plot
minerals (Spears and Kanaris-Sotiriou, 1976). above the trace element versus Al2O3 minimum curves in x–y plots
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 73

Fig. 5. Relationship between Al2O3 and selected trace elements, Fe2O3, and Corg in the KCF. The Th:Al2O3 ratio in samples from all sections except OPI-2/1 is close to that of the world
shale average, or WSA (Gromet et al., 1984), and the single trend extrapolates to the origin. We can offer no explanation for the seemingly high Al2O3 (low Th) values in OPI-2/1. The
Fe2O3 values in all units except OPI 2 scatter above and below the curve for the WSA, suggesting a lack of Fe enrichment, on average. The trace elements other than Th and Fe2O3 are
uniformly enriched above a lithogenous contribution, represented by their displacement above a curve drawn through the samples with the lowest trace-element:Al2O3 ratios. The
slopes of these minimum curves in frames D–H closely approach trace-element:Al2O3 ratios for the different trace elements in the WSA, shown by asterisks. The displacement of
samples above the curves in these frames is taken as the seawater-derived marine fraction of trace elements (see text). The relationships between Al2O3 and Corg (frame C) in samples
of the separate units within the Cleveland Basin exhibit negative trends with the omission of a single analysis from OPI-1. Correlation coefficients and p values of the linear
regressions for the different units are shown in the legend in parentheses. Samples from the Boulonnais Basin site exhibit a positive trend, but the correlation is weak and the p value
is large. Nonetheless, the samples are distinct from those of units from the Cleveland Basin. The shaded areas in frames F and G enclose samples from Argiles de Chatillon.
74 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

Westerlund, 1988, 1991), its rate of deposition can be used to estimate


the flux of Corg settling out of the photic zone. This interpretation is
based on the observation that Ni is intimately involved in biogeo-
chemical cycling in the ocean, for which its mean concentration
measured in plankton of ca. 18 ppm (Collier and Edmond, 1983) is in
close agreement with the value calculated (Piper and Dean, 2002)
from its depth profile in the ocean (Sclater et al., 1976; Whitfield,
2001; Cotté-Krief et al., 2002; Fig. 4). The Ni:P ratio measured in
organic matter is 2.1 × 10− 3 g g− 1, whereas the ratio obtained from
the increase in the concentration of dissolved Ni and P from the photic
zone to below the main pycnocline in the Pacific Ocean, in response to
remineralization of organic matter, is 2.7 × 10− 3 g g− 1. In the North
Atlantic, the profiles of Ni and HPO2− 4 in the upper ca. 500 m of the
water column yield a ratio that is close to 2 × 10− 3 g g− 1 (Cotté-Krief
et al., 2002). By contrast, both the addition of Ni to the O2-depleted
Fig. 6. Relationship between Hf and Zr in samples from the Boulonnais Basin site in part of the water column of anoxic basins (e.g., Saanich Inlet,
northern France. Their ratio is similar to the mean for WSA, the North American Shale Framvaren Fjord, the Cariaco Basin (Fig. 7) and the Black Sea) via
Composite (NASC), the upper continental crust (UC), and the lower continental crust remineralization of organic matter and its removal by inorganic
(CC). Data for the standards are from Wedepohl et al. (1969–1978), Gromet et al.
(1984), and Wedepohl (1995).
processes appear to be insignificant (Jacobs et al., 1985, 1987; François,
1988; Haraldsson and Westerlund, 1991) because its concentration is
uniform with depth. Thus, the accumulation of the seawater-derived
fraction of Ni in sediments that accumulate under O2 depleted
(Fig. 5) identify the marine fraction of trace elements (Piper et al., bottom-water conditions provides an indirect estimate of Corg settling
2007) that are contributed by biogenous debris (biogenous fraction) out of the photic zone and its deposition on the seafloor prior to burial
and precipitates and adsorbed phases (hydrogenous fraction) on a and post-depositional oxidation of the deposited organic matter.
sample-by-sample basis. Concentrations of individual trace elements In contrast to the behavior of Ni, the biogenous contribution of Mo
and inter-element ratios within the marine fraction alone may and V to sediments is minor. Concentrations in phytoplankton average
identify relative levels of primary productivity and bottom-water less than 2 to 3 ppm (Brumsack, 1986) and their dissolved con-
redox conditions. As an example, the Ni:Al2O3 ratio in the bulk centrations exhibit minor enrichments with depth below the photic
sediment identifies the marine input of Ni to the seafloor of an anoxic zone (Fig. 4). Vanadium is reduced to the strongly adsorbed VO(OH)− 3
basin (Piper and Dean, 2002). From our understanding of its under moderate suboxic conditions (Emerson and Huested, 1991).
geochemistry in seawater (Jacobs et al., 1985, 1987; Haraldsson and Conversely, high dissolved sulfide concentrations are required to

Fig. 7. Depth profiles of selected trace elements and nutrients in the Cariaco Basin. The profiles are similar in other anoxic basins, but absolute concentrations vary slightly. For
example, Mo in the Black Sea has its highest concentration of ca. 75 nmol L− 1 in the very surface water and decreases more or less uniformly to ca. 5 nmol L− 1 at 300 m depth, below
which it maintains a uniformly low concentration (Emerson and Huested, 1991). The particulate Mn maximum in the Black Sea is also in the suboxic zone but at a much shallower
depth; otherwise, its profile and that of dissolved Mn (Brewer and Spencer, 1974) are virtually the same as in the Cariaco Basin. The Fe profiles in the two basins also are similar. In the
Cariaco Basin, the particulate Fe distribution is complicated by a high lithogenous contribution (Bacon et al., 1980), particularly in the near-surface water. The data for Mo and V are
from Emerson and Huested (1991); data for Cd, Ni, dissolved Fe, and dissolved Mn are from Jacobs et al. (1987); the data for particulate Mn, HPO2− 4 , and H4SiO4 are from Richards
(1975). The particulate Fe curve represents a combining of data from Richards (1975) and Bacon et al. (1980). The Ni curve represents a three-floating-point average. The Ni data of
Haraldsson and Westerlund (1988) for the Black Sea show that it is not significantly depleted in the deep water, similar to its distribution shown here for the Cariaco Basin. By
contrast, Landing and Lewis (1991) report a 40% lowering of Ni in the bottom water of the Black Sea. Profiles of fixed-N in both basins are similar to those shown in Fig. 3. The color
coding is given in Fig. 3. As chemical interfaces migrate as much as 100 m seasonally (Thunell, personal communication, 2007), some adjustment of sample depths has been
attempted based on the depth of the suboxic–anoxic interface.
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 75

initiate efficient conversion of soluble molybdate to thiomolybdate section thicknesses between dated horizons are the presence of
species, which are then fixed as S complexes (Helz et al., 1996). These laterally supplied sediment in the sedimentary section (focusing) and,
reactions lead to the removal of V and Mo from O2-depleted seawater conversely, the removal of material by sea-floor erosion (winnowing).
(Emerson and Huested, 1991) and accumulation in sea-floor sedi- These problems have been highlighted in Late Quaternary sediments
ments under slightly different bottom-water redox conditions, V of the mid-latitude Atlantic (François et al., 1990), the Southern
under suboxic and anoxic bottom-water conditions and Mo only Ocean (Frank et al., 1995), and the equatorial Pacific (Marcantonio
under fully anoxic conditions. The V:Mo ratio in the marine-sediment et al., 2001; Loubere et al., 2004; Anderson et al., 2006; Kienast et al.,
fraction will then distinguish between these two redox conditions 2007). For Holocene and Late Pleistocene sediments, corrections for
of bottom waters (Piper and Isaacs, 1994, 1996; Piper et al., 2007). these effects have been developed using the 230Th-normalization
Under fully anoxic conditions, the V:Mo ratio in the sediment should technique (François et al., 2004). Because of the short half-life of
230
approach its ratio in seawater itself, ca. 1.8:10. The weak biogenous Th, this approach is only applicable for sediments younger than
signal of both metals and their different efficiencies of removal from ca. 350,000 years.
anoxic seawaters (Emerson and Huested, 1991; Piper and Dean, 2002)
will change this ratio in the sediment slightly, but should not increase 2.5.4. Limitations of modeling trace-element ratios and accumulation rates
it significantly above ca. 2. Under suboxic bottom waters, i.e., where Estimates of the concentrations of trace elements in the two
SO2−4 reduction is restricted to the sediment pore water, V:Mo ratios in marine fractions of sediment become problematic (1) when the
the sediment will be elevated above their seawater ratios. Sediments marine contribution of an individual trace element approaches the
that accumulated under bottom-water conditions that varied through uncertainty of the analytical technique and (2) when the composition
time from suboxic to anoxic should also exhibit elevated marine V:Mo of the lithogenous fraction of the sediment, represented by the slopes
ratios. A sampling scheme for studying sedimentary rocks, for which a of trace elements versus Al2O3 plots (Fig. 5), fails to identify a single
centimeter section can represent as much as several hundred years, composition that is similar to the various rock standards. In the case
imposes on an individual sample the probability that it may have considered here, the adoption of a single composition for this fraction
recorded multiple conditions of bottom-water redox. Conversely, an in the KCF is supported by the close similarity of the slopes of the trace
extended sediment section that exhibits a strongly uniform V:Mo ratio element versus Al2O3 curves to their ratios in the crustal standards.
would be indicative of a stable water column. Crusius et al. (1996) The use of bulk-sediment concentrations and concentration ratios,
showed that the accumulation rates of Re and U driven by their without adjusting for contributions by siliciclastic components,
diffusion across the benthic boundary (Anderson et al., 1989) have a severely compromises the modeling of elemental fluxes. This ap-
relationship with Mo in sediments that is similar to the observed V proach limits the sensitivity and accuracy of identifying and inter-
and Mo relationship in the sediments. preting the marine sources of diagnostic elements. For example, the
Tribovillard et al. (2008) have cautioned against the use of Mo bulk-sediment V:Mo ratio of the Black Sea surface sediment is 7
enrichment as a proxy for bottom-water anoxia in ancient sediments (Volkov and Fomina, 1974; Calvert, 1990), largely reflecting the much
in the absence of enrichments of other redox-sensitive trace elements, higher V concentration in lithogenous siliciclastic detritus (100 to
e.g., Re, U, and V. They note that Mo may be taken up by framboidal 150 ppm) than Mo (2 to 3 ppm). Given the seawater concentrations of
pyrite during early diagenesis under oxic to suboxic bottom-water V (1.8 ppb) and Mo (10 ppb) and the efficiency of their removal from
conditions. This mechanism of Mo enrichment may result in an over- the Black Sea anoxic waters of 75% and 95%, respectively (Emerson
estimation of the importance of Mo enrichment as solely indicative of and Huested, 1991), their seawater-derived V:Mo ratio is unlikely to
anoxic bottom-water conditions. Zheng et al. (2000) and McManus be greater than one. The ratio of their marine fractions in surface
et al. (2006) have further reported Mo accumulation under less than sediment from the Black Sea is indeed less than 1.5 (see Section 3.2.4).
anoxic conditions. Thus, bottom waters may have been less reducing, The V:Mo ratio in surface sediment from the Cariaco Basin is 0.65
i.e., suboxic, when Mo concentrations alone indicate the bottom water (Piper and Dean, 2002). A significantly higher ratio in the marine
was anoxic. In such sedimentary rocks, framboidal pyrite should be an fraction of a number of sedimentary deposits, e.g., the Quaternary
important component of the sediment and host of Mo that had a sediment from the Japan Sea (Piper and Isaacs, 1996), the Miocene
seawater source (Tribovillard et al., 2008). Monterey Formation of California (Piper and Isaacs, 1994), the Permian
Phosphoria Formation of the northwest United States (Piper, 2001),
2.5.3. Trace-element accumulation rates and the Devonian Ohio Black Shale of the eastern United States
Although the trace-element ratios discussed here identify bottom- (Perkins et al., 2008), must reflect deposition under dominantly
water redox conditions of deposition, they do not provide information suboxic conditions.
on rates of hydrographic, biological, and chemical processes that The bulk Ni:Co ratio, which has been used to identify anoxic water-
contribute to an understanding of the dynamics of the sedimentary column conditions (Jones and Manning, 1994; Rimmer et al., 2004),
environment e.g. rates of nutrient and trace-element supply to a basin, has even more severe limitations for deducing the redox status of
primary productivity, upwelling, and advection of bottom water and sediments. Nickel has both a lithogenous and a biogenous source,
its residence time (RT) in a basin. The accumulation rates of individual whereas Co has primarily a lithogenous source (Brewer and Spencer,
trace elements alone allow these aspects of deposition to be addressed 1974; Jacobs et al., 1985, 1987). Therefore, the uncorrected bulk-
and for that sediment accumulation rates are required. Bulk-sediment sediment ratio provides no information about bottom-water anoxia
accumulation rates can be estimated by assuming that the linear and less precise information about the possible enrichment of Ni by
thickness of a sedimentary section between two horizons of known the biogenous fraction than given by the Ni:Al2O3 ratio. We should add
age represents the flux of sedimentary material settling through the that Co does have a hydrogenous source under oxic conditions (see
water column during that time interval (Eq. (1)), as opposed to the Section 3.2).
lateral transport of sediment. The accumulation rates of individual The V/(V + Ni) ratio, or some variant of the ratio, in bulk sediment
trace elements then provide insight into (1) the level of organic has also been used to identify anoxic conditions in the geologic record.
production from the accumulation rate of trace nutrients and (2) the Enrichments of both metals were ascribed by Lewan and Maynard
hydrography, or the circulation and the oxygenation of the waters (1982) to the metallation of porphyrins, which are themselves
from the accumulation rates of redox-sensitive trace metals. enriched in fine-grained shallow-water sediments that have anoxic
Linear sedimentation rates are commonly assumed in calculations sediment pore waters, but were not necessarily deposited from anoxic
of trace-element fluxes to the seafloor. Important caveats to the bottom waters. Thus, the use of the bulk-sediment V/(V + Ni) ratio to
modeling procedure employing sedimentation rates from linear identify anoxic bottom-water conditions of sedimentation, as opposed
76 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

to sediment pore waters, is an inappropriate extension of the findings environments are further distinguished by being more or less per-
of Lewan and Maynard (1982). manently suboxic, seldom anoxic (Dugdale et al., 1977). As discussed
These results show that the estimate of the strictly marine previously, the seafloor beneath some OMZs can be densely inhabited
contributions of trace metals to a given deposit yields a much more by macrofauna, owing to a large flux of O2 over the seafloor delivered
reliable reconstruction of past geochemical properties of marine by a strong bottom current, despite having near zero O2 concentration
waters, and that bulk-sediment compositions and ratios are less (Arthur et al., 1998).
informative in this respect. An exception to this generalization occurs within the OMZ
between 200 and 800 m depth in the Gulf of California and off
3. Modern marine environments northwestern Mexican (Calvert, 1964; Ganeshram and Pedersen,
1998). Finely laminated sediment is currently accumulating on these
Studies of the geochemistry of the sediments in several modern continental slopes under suboxic conditions. Similar to a true basin
anoxic marine basins have provided insight into the accumulation of setting, the gulfs are partially separated from the open ocean. In the
those trace elements that behave as proxies for the delivery of case of the Gulf of California, it is isolated from the California Current,
biogenic materials to the basin floor (biogenous sediment fraction), the eastern boundary current of the Northeast Pacific Ocean, but it is
as proxies of water-column redox (hydrogenous sediment fraction), without a sill that would otherwise limit water exchange with the
and as proxies of the relatively non-reactive siliciclastic sediment Pacific. A second possible analogue is the inner shelf off Namibia,
(lithogenous sediment fraction). The rate of accumulation of the where organic-rich, highly sulfidic diatomaceous muds are currently
biogenous fraction reflects rates of nutrient supply to the photic zone, accumulating on a 100 m deep platform beneath the highly
i.e., primary productivity, by upwelling and/or mixing, whereas the productive Benguela Current (Calvert and Price, 1983). Bottom
rate of accumulation of the hydrogenous fraction tracks changes in waters here are seasonally suboxic due to the high settling flux of
bottom-water advection and redox. The rate of accumulation of the planktonic debris (Calvert and Price, 1971; Bailey, 1991; Tyrrell and
lithogenous fraction provides an estimate of the dilution of the Lucas, 2002). Occasional fish kills (Copenhagen, 1953) are caused by
marine fractions by terrigenous debris. We review some of the results the intermittent presence of dissolved sulfide in the waters (Brüchert
of studies of two anoxic basins—the Cariaco Basin and the Black Sea— et al., 2003). In this view, a black shale basin may not necessarily
to illustrate the processes that govern sediment compositions in have been a bathymetric feature, but a site of deposition where
these two environments. The calculations that follow provide a bottom circulation was restricted by shelf breadth and only partially
quantitative illustration of trace-element accumulation rates on the isolated from the open ocean. This usage follows that developed from
seafloor of these two basins that are consistent with their different the extensive work on the modern sedimentary environments of the
hydrographies and water-column geochemistries. The results then Gulf of Mexico and the Gulf of California (Shepard et al., 1960; van
serve as a guide for our interpretation of the environment of Andel and Shor, 1964). A basin, or shelf depth less than ca. 300–
deposition of the KCF. 400 m, would further allow the most intensely negative redox
The seafloor in the Cariaco Basin and Black Sea reach depths of conditions to occur at the seafloor rather than at an intermediate
1400 m and 2200 m, respectively; sill depths are 140 m and 40 m, depth of an OMZ.
respectively. Water-column redox profiles in the two are similar The extrapolation of relationships between the marine chemistry
(Fig. 3). They have a surface mixed, oxic layer several tens of meters and hydrography of modern basins to those properties of ancient basins,
thick. In the Black Sea, the thickness of the surface layer has decreased based on trace-element accumulation rates, are certainly limited by
in recent years, owing to anthropogenic inputs of nutrients (Polat and uncertainties related to variations through time of: (1) the trace-
Tuğrul, 1995, and references in Özsoy and Ünlüata, 1997). Both basins element geochemistry in seawater (Algeo, 2004): (2) the uptake of trace
have an intermediate suboxic layer, in which bacterial respiration nutrients by plankton (Quigg et al., 2003): (3) diagenesis of sediments
lowers the concentration of fixed-N to zero. In the Black Sea, the following deposition and burial: (4) weathering following uplift and
suboxic layer is only about 25 to 75 m thick, whereas it is as much as exposure: and (5) uncertainties in estimates of the trace-element
150 m in the Cariaco Basin (Thunell et al., 2000; Goñi et al., 2003). The marine faction of sediment using a single model for the lithogenous
bottom water in both basins is sulfidic. fraction. For these reasons, the following calculations that elucidate the
The basins are useful modern analogues for ancient black shales, oceanographic dynamics of modern basins in the short term can only be
although we do not imply by this that barred basins are the favored applied to time-averaged conditions of an ancient basin.
paleobathymetric model for basins in which black shales have
accumulated. Sediment texture of both deposits reflects the partial 3.1. The Cariaco Basin
isolation from the ocean that renders their sea-floor environments
tranquil. Their sediments are fine-grained and finely laminated. Laminae The Cariaco Basin is a relatively small depression on the con-
in sediment recovered from a 900 m deep saddle in the Cariaco Basin, tinental shelf of Venezuela, consisting of two 1400 m-deep sub-basins
which separates two 1400 m deep sub-basins, represent annual varves separated by a 900 m deep saddle. It is partially isolated from the
(Hughen et al., 1996). The sediment texture of the KCF (Wignall, 1990) Caribbean by a 145 m deep sill. During the last glacial maximum
and of other black shales in the geologic record is also commonly (LGM), the basin was nearly isolated from the Caribbean Sea. A rising
laminated, suggesting low-energy bottom-water conditions. sea level, beginning about 18 ka (Fairbanks, 1989), imposed a complex
This property contrasts sharply with environments on open signal of temporal changes on rates of upwelling, bottom-water
continental shelves, where oxygen-depleted seawater in the OMZ advection, and the flux and composition of lithogenous debris into the
intersects the shelf at depths between roughly 100 and 800 m, basin (Peterson et al., 1991; Yarincik et al., 2000). These changes are
(Wyrtki, 1962, 1963; Codispoti, 1980, 1989; Kuypers et al., 2005). For now reflected in the lithology of the sediment. In a piston core from
example, sediments on the Peru shelf show strong textural and the 900 m deep saddle (Peterson et al., 1991, 1995), the sediment is
geochemical evidence of suspension and redeposition by the under- bioturbated below approximately 600 cmbsf (centimeters below the
current beneath the Peru Current (Kim and Burnett, 1988; Arthur and seafloor). It is distinctly laminated above this unit to about 300 cmbsf,
Dean, in press) that sweeps over the shelf. This property is a and less distinctly and more thinly laminated from 300 cmbsf to the
prominent feature of other open ocean, continental shelf deposits sediment surface. The laminae consist of alternating biogenous
that exhibit high primary productivity (van Andel and Calvert, 1971; (carbonate and/or diatomaceous silica) and clay-rich layers, each
Arthur et al., 1998), in marked contrast to the texture of sediments lamina couplet representing an annual varve (Hughen et al., 1998).
accumulating in the tranquil environment of silled basins. The shelf Peterson et al. (1991) interpreted the lithologic boundary between the
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 77

bioturbated and the laminated units, dated at 14.8 ka (Lin et al., 1997; equivalent to oxidation of the Corg, is given by their reduction in the
Hughen et al., 1998), as recording the onset of elevated primary lowermost 1100 m of the water column:
productivity in the photic zone and of SO2−4 reducing conditions in the
bottom water. These properties of the water column continue to the Oxic respiration : 10CH2 O + 10O2 + 10CaCO3 = 20HCO3− + 10Ca2 +
present (Richards, 1975). Corg oxidation = reduction of oxidantðsÞ
Partial isolation of the Cariaco Basin and a high settling flux of Corg
to depth (Thunell et al., 2000; Goñi et al., 2003) maintain anoxic ½
= 0:185 mol O2 m
−3
× 0:4 × 1100 m
conditions in the bottom water. Oxygen concentrations decrease from 
× 1:0 mol Corg mol − 1 O2 = 82 mol Corg m − 2 :
a maximum of ca. 210 µM in near-surface waters (Fig. 3) to zero at ca. ð3Þ
300 to 350 m depth (Richards, 1975; Scranton et al., 2001), where H2S
increases with depth, reaching a concentration of ca. 40 µM in the Suboxic respiration : 10CH2 O + 8NO3− + 2CaCO3
abyssal waters (Fanning and Pilson, 1972; Emerson and Huested, 1991; − 2 +
= 12HCO3 + 4N2 + 4H2 O + 2Ca Corg oxidation
Scranton et al., 2001). The water column at intermediate depth (150 to h 
− −3
= 0:005 mol NO3 m × 0:70
300 m) is suboxic, or denitrifying. Fixed-N, mostly as NO− 3 , generally is
 
very low in the near-surface waters, but increases to ca. 12 µM at
200 m due to nitrification. Below this depth, it decreases to zero at 300
× 1100 m × 1:25 mol Corg mol − 1 NO3− 
to 350 m. The other important forms of fixed-N (NH+ −
= 5 mol Corg m − 2 : ð4Þ
4 and NO2 ) are
also forced to zero at this same depth (Fig. 3) by anammox reactions
(Schmidt et al., 2002; Dalsgaard et al., 2003; Kuypers et al., 2003,
Anoxic respiration : 10CH2 O + 5SO24 − + 5CaCO3
2005), by some variant of the following reaction:
= 15HCO3− + 5HS − + 5Ca2 + Corg oxidation
h   i
þ − −
3:33NH4 + 2NO2 + 0:66HCO3 + 0:33CaCO3 = 0:040 cmol SO24 − m − 3 ×1100m × 2 mol Corg mol − 1 SO24 −

= 2:66N2 + CH2 O + 0:33Ca


2 +
+ 6H2 O: ð2Þ = 88 mol Corg m − 2 : ð5Þ

In Eq. (4), the value 0.70 adjusts for the lower oxidative potential of
Below this depth ammonium concentrations increase, reaching
NO−3 owing to anammox reactions (Eq. (2)); 1.25 is the ratio of Corg
ca. 15 µM in the deep waters. Phosphate and silicic acid are also in low
oxidized versus NO− 3 reduced. In Eq. (5), the 0.04 mol m
−3
of SO2−
4
concentration in the near-surface waters (Fig. 7), but increase
reduced is simply the concentration of H2S in the bottom water; the
smoothly with depth below the pycnocline to maximum values in
2 mol Corg mol− 1 SO2−
4 is the molar ratio of Corg oxidized versus SO4
2−
the abyssal waters of 3 and 70 µM, respectively (Fanning and Pilson,
reduced. If we ignore the consumption of SO2− 4 by CH 4 oxidation, the
1972; Richards, 1975).
calculated rate of Corg oxidation in the bottom water of 1.34 mol Corg
Primary productivity is highly seasonal but averages 550 g Corg
m− 2 yr− 1 (70% of 1.92, or annual consumption rate of oxidants) gives
m− 2 yr− 1 (Varela et al., 1997; Goñi et al., 2003). It is supported by
the following RT:
the input of nutrients in sub-thermocline waters from the Caribbean
Sea (Richards, 1975). Goñi et al. (2003), Müller-Karger et al. (2001), Bottom  water RT
and Thunell et al. (2000) have shown that primary productivity is −2 −2 −1
highest between December and April, when the Intertropical = ð82 + 5 + 88Þmol Corg m  1:34 mol Corg m yr
Convergence Zone (ITCZ) lies south of the equator and the strong = 130 years; ð6Þ
northeasterly trade winds create intense upwelling along the coast-
line. Water entering from the Caribbean across the sill at approxi- approximating the 100 years calculated by Deuser (1973). The slightly
mately 100 to 200 m depth (Müller-Karger et al., 2001) upwells into greater RT calculated here may represent uncertainty in the O2
the photic zone when the near-surface thermocline is weakly consumed in the oxic and suboxic layer extending to 350 m depth
developed and the 22 °C isotherm penetrates to the surface. During versus 275 m (the depth of the sediment traps), a variable H2S
the summer and fall a strong thermocline develops at 50 m and concentration in the bottom water (Fanning and Pilson, 1972), and the
upwelling and primary productivity are depressed. reduction of NO− +
3 to NH4 rather than to N2 (Eq. (4)). Despite the
permanent anoxic conditions in this shallow-silled basin, the bottom-
3.1.1. Oxidant balance in and residence time of bottom water water RT reflects a rather vigorous renewal of bottom water compared
The mean flux of Corg settling into the anoxic bottom water is with the renewal of bottom water in the Black Sea (see Section 3.2.3).
1.92 mol m− 2 yr− 1 (Thunell et al., 2000; Goñi et al., 2003); 30% is
buried on the seafloor and 70% is oxidized in the water column and at 3.1.2. Trace-metal budgets
the benthic boundary. As essentially all of the Corg has a marine origin Jacobs et al. (1987) suggested that the bottom water in the Cariaco
(Wakeham and Ertel, 1988), the carbon oxidized in the bottom water Basin enters at the same approximate depth as the water upwelling
(1.34 mol m− 2 yr− 1) can be used to estimate the bottom-water RT into the photic zone. This may be explained by the importation of
from the consumption of oxidants. We ignore the flux of H2S into the bottom water largely in the summer and fall below the strong
suboxic water and its subsequent oxidation. The concentrations of O2 thermocline that develops at 50 m depth (Müller-Karger et al., 2001;
and NO− 3 in the water entering at ca. 200 m depth (Müller-Karger Goñi et al., 2003). The strong thermocline and the depth of imported
et al., 2001) are taken as their concentrations in the Caribbean at this water at this time of year limit upwelling and mixing into the
same depth (Morrison and Smith, 1990), namely 185 μM O2 and 5 μM shallower depths, possibly enhancing water intrusion to depth. In the
NO− 3 . The O2 is adjusted downward by 60%, the amount estimated as winter and spring, much of the water imported probably mixes up
consumed in the suboxic layer, based on the difference between the into the photic zone, owing to the collapse of the thermocline during
10% of the Corg fixed by photosynthesis and assumed to settle out of these seasons (Goñi et al., 2003). Such a seasonal difference in the
the photic zone and the 4.2% of the primary production settling to import of surface water and bottom water from a single depth in the
275 m. The 10% used here is identical to the global average of Corg Caribbean accounts for the similarity in salinity and temperature of
settling out of the photic zone (Suess, 1980; Berger et al., 1988), which the bottom water of the basin and of the Caribbean Sea at 100 to 200 m
in turn is close to the value for the Black Sea (Karl and Knauer, 1991). depth (Richards, 1975), as well as the similarity in Mo and V
The sequential consumption of the three oxidants in the bottom water, concentrations in the Cariaco Basin water column above the suboxic
78 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

layer (Emerson and Huested, 1991; Fig. 7) and their concentrations in section to be calculated from bulk-sediment accumulation rates,
the open ocean. independently of their accumulation within the lithogenous fraction
First, consider the budget of trace nutrients in the water column (Piper and Dean, 2002). The Ni rate can then be compared to its
and sediment, focusing on Ni. It exhibits little or no depletion in the settling rate out of the photic zone incorporated in particulate organic
photic zone (Fig. 7), yet Collier and Edmond (1983) have measured a matter, again assuming a Ni concentration in plankton of 18 ppm
concentration in plankton of ca. 18 ppm. Its depth profile in the (Collier and Edmond, 1983) and a concentration of Corg in organic
Venezuela Basin, relative to that of HPO2− 4 , yields a concentration of matter of 36% (Redfield et al., 1963). For V and Mo, their accumulation
25 ppm in organic matter, for which the organic matter is assumed to rates are compared to their rates of removal from the bottom water
have Redfield stoichiometry (Redfield et al., 1963), i.e., 36% Corg. A as the hydrogenous fraction, which are given by their level of removal
concentration of 25 ppm is consistent with its measured concentra- from the bottom water of 65% and 15%, respectively, i.e., their
tion in average plankton of 18 ppm. Although we are not able to concentration in the surface oxic water versus their concentration in
calculate its rate of input into the photic zone, owing to a weakly the bottom water (Emerson and Huested, 1991), and the bottom-
constrained upwelling rate, we can compare its availability in the water RT.
upwelling water to its uptake by phytoplankton, relative to that of The deposition of marine Ni is given by the settling flux of Corg out

HPO2−4 , or NO3 . As HPO4
2−
and NO− 3 are completely removed in the of the photic zone (10% of primary productivity) and the mean Ni
photic zone on a seasonal basis, both are limiting nutrients to primary (Collier and Edmond, 1983) and C org (Redfield et al., 1963)
productivity. The Ni:HPO2−4 ratio in the water fluxing into the basin at concentrations in phytoplankton.
ca. 200 m depth is 0.12:190 μg/µg (Jacobs et al., 1987), or 0.00063. The  
same ratio in plankton is 18:0.26 × 10− 6, or 0.000067. Thus, only about Ni deposition rate = 550 g Corg m
−2
yr
−1
× 0:10
10% of the Ni made available through upwelling is utilized by  
× 18 cg Ni  0:36 g Corg
phytoplankton, in agreement with only a slight depletion in the
−2 −1
photic zone (Fig. 7). The same calculation for Cd reveals that 90% of = 2750 cg Ni m yr ; ð7Þ
the Cd upwelling into the photic zone is taken up by phytoplankton,
accounting for its near complete depletion in the photic zone. This rate can be compared with its accumulation rate on the seafloor,
Marine Ni is the preferred proxy for this biogenic fraction in the the product of the mean bulk-sediment accumulation rate and the
sediment, as it has only the one marine source, biogenic detritus. That marine Ni concentration in the sediment:
there is little input of Ni into the sediment from the anoxic water
−2 −1 −1
column is shown by its uniform concentration in the bottom water of Ni accumulation rate = 138 g m yr × 19 cg Ni g
this and other anoxic basins (Kremling, 1983; Jacobs et al., 1985; = 2620 cg Ni m
−2
yr
−1
: ð8Þ
Haraldsson and Westerlund, 1991), although its removal under
sulfidic conditions as the sparingly soluble sulfide would have been The accumulation rate of Corg itself is 6.9 g m− 2 yr− 1 (Piper and Dean,
predicted. Conversely, the seawater profile of Cd reflects its uptake in 2002), which has a concentration in the upper few meters of the
the photic zone by phytoplankton, remineralization below the photic sediment of 5 wt.% (±1%). This represents only 1.25% of primary
zone of the organic matter with a Cd concentration of 12 ppm productivity, suggesting no enhancement of Corg preservation under
(Brumsack, 1986; Collier and Edmond, 1983), and its additional anoxic bottom-water conditions, as deduced by Thunell et al. (2000).
removal from the bottom water as a hydrogenous fraction. Copper and For marine Mo, its deposition from the bottom water between 300
Zn have similar dual pathways in the ocean to that of Cd. and 900 m depth onto the ca. 900 m deep saddle, represented by the
Molybdenum has a concentration of ca. 10.5 ppb in the uppermost removal of 15% of its inventory in the bottom water (Fig. 7), is:
300 m of the water column (Fig. 7), similar to its concentration in the
open ocean (Collier, 1985) and, presumably, in the Caribbean. It shows Mo deposition rate
   
a step decrease (Fig. 7) of 15% at 300 m depth (Emerson and Huested, = 10:5 cg Mo L
−1 5
× :15 × 6 × 10 L m
−2
 115 yr
1991) and remains lower throughout the bottom water than its −2 −1
concentration at the surface. Vanadium shows a similar profile to that = 8300 cg Mo m yr ; ð9Þ
of Mo, except that it decreases smoothly through the suboxic zone,
reaching a concentration in the anoxic bottom water of 35% of its at a mean bottom-water RT of 115 years. This compares closely with
surface value (Emerson and Huested, 1991). As both metals are weakly its accumulation rate in the sediment:
bioreactive and strongly redox reactive, their enrichment in the −2 −1 −1
sediment is largely as a hydrogenous fraction; they should serve as Mo accumulation rate = 138 g m yr × 78 cg Mo g
ideal proxies of bottom-water anoxia. = 10; 600 cg Mo m
−2
yr
−1
: ð10Þ
The accumulation of the third sediment fraction, the lithogenous
fraction along with its complement of trace elements, is seasonally Similarly, the deposition of V from bottom water is 9300 μg m− 2 yr− 1,
variable, owing to variable stream runoff from the coast of Venezuela and its accumulation rate as a hydrogenous fraction is 7200 μg m− 2 yr− 1.
that leads to an increased flux of lithogenous sediment into the basin Although sediment-trap studies in the Black Sea (Crusius et al., 1996),
that is out of phase with primary productivity and the flux of Corg to Saanich Inlet (François, 1988), and the Cariaco Basin (Calvert, Piper, and
the seafloor (Goñi et al., 2003). The higher production and settling Thunell, in preparation) reveal that Mo removal from the bottom water is
flux of biogenous components (CaCO3, opal, and organic matter) in restricted to the benthic boundary layer in these anoxic basins (elevated
winter and spring and the higher delivery of lithogenous (siliciclastic) concentrations are not captured in sediment traps), its accumulation rate
material in summer and fall leads to the deposition of the on the seafloor (Eq. (10)) equals its removal rate from the water column
compositionally distinct laminae in the sediment on the centrally- (Eq. (9)), which is calculated from its depletion in the entire 600 m of
located saddle (Hughen et al., 1996). Trace elements and major- anoxic water above the core site (Piper and Dean, 2002). The
element oxides that can serve as proxies of the lithogenous fraction hydrogenous V fraction and the biogenous Ni fraction are both captured
include the rare-earth elements, Th, Ga, Co, Li, Rb, Cs, K2O, Fe2O3, and by sediment traps located within the water column (Calvert, Piper, and
of course Al2O3 (Piper and Dean, 2002). Thunell, in prep.). In other words, their settling fluxes equal their
The normative modeling scheme described earlier allows the deposition rates. The close agreement between the rates of accumulation
accumulation rates of the marine fraction of the trace elements, on the seafloor of the marine fractions of Mo, Ni, and V in the sediment
specifically Mo, Ni, and V, in the uppermost 50 cm of the sediment with their rates of delivery to the seafloor in this modern basin, based on
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 79

the level of primary productivity (Ni) and their distributions in, and RT of, delivered by the Danube (ca. 50%), the Dnestr and Dnepr (together ca.
bottom waters (Mo and V) supports the modeling of trace-element 40%), and smaller rivers (10%). These inputs of fresh water and the
accumulation rates in sediments to provide estimates of the levels of difference between evaporation (353 km3 yr− 1) and precipitation
primary productivity in the photic zone and the exchange rate of bottom (300 km3 yr− 1) maintain mass balance of salt and water, driving the
water with the Caribbean. estuarine circulation between the Black Sea and Sea of Marmara.

3.2. The Black Sea 3.2.1. Primary productivity


Primary productivity in the Black Sea is ca. 210 g Corg m− 2 yr− 1
The Black Sea was isolated from the Mediterranean during the last (Sorokin, 1983; Karl and Knauer, 1991). Production is seasonally
glacial maximum, when sea level stood ca. 130 m lower than its variable, diatoms blooming in February–May and coccoliths blooming
present level (Caspers, 1957; Strakhov, 1971; Degens and Ross, 1972; in June–October (Sorokin, 1983; Honjo et al., 1987). Lithogenous
Fairbanks, 1989). Post-glacial eustatic sea-level rise connected the particle fluxes are highest in November–May, the seasonal change in
Mediterranean with the Black Lake when global sea level reached the the composition of the settling material leading to the formation of
level of the Bosporus sill, initiating the influx of saline waters and compositionally distinct laminae in the bottom sediment (Müller and
transforming the Black Sea into the modern, intermediate-salinity, Blaschke, 1969; Müller and Stoffers, 1974). Although these alternating
estuarine basin. Hydraulic modeling has suggested that sea water biogenous and lithogenous laminae may constitute annual varves
invasion began roughly 1000 years after sea-level reached the level of (Arthur et al., 1994), reliably counting them in abyssal basin cores is
the Bosporus sill at 40 m depth (Lane-Serff et al., 1997), a delay somewhat tenuous (Crusius and Anderson, 1992).
required to set up a sufficient hydraulic head above the level of out- Primary productivity is sustained by the import of nutrients into
flowing fresh water from the Black Sea so that two-layer flow could be the photic zone from approximately 50 m depth. The NO− 3
established. The accumulation of brackish and marine deposits, concentration at this depth, which we take as the limiting nutrient,
beginning with the Black Sea sapropel, began at around 7500 to is 8 μM (Codispoti et al., 1991). The average fluvial NO− 3 input is
7800 years BP (Jones and Gagnon, 1994; Arthur and Dean, 1999). obtained from the product of the mean river concentration prior to the
Today the narrow Bosporus Strait connects the basin to the eastern last few decades (Maybeck, 1982) and the river flux:
Mediterranean and the Sea of Marmara. It is the largest permanently
− −1 3 −3 11 3 −1
anoxic basin in the marine realm today. It has been variously described Riverine NO3 flux = 15 cmol L × 10 L m × 3:52 × 10 m yr
as “the type euxine” (Byers, 1977), the “type anoxic basin” (Glenn and ¼5:3 × 10
15
cmol NO3

yr
−1
: ð11Þ
Arthur, 1984), “the definitive anoxic basin” (Murray, 1991), and “a freak
of palaeogeography” (Tyson, 2005). Its isolation from the open ocean Anthropogenic inputs of NO− 3 and other nutrients by the rivers
and the permanently anoxic state of its bottom waters render the Black entering the Black Sea have increased this value possibly by as much
Sea a highly informative geochemical model of ancient basins. as 10-fold in the last few decades (Maybeck, 1982; Oguz, 2005), but
The salinity contrast between surface and deep waters of the Black we ignore this perturbation here. An estimate of the NO−3 input from
Sea induces a strong density stratification, maintaining anoxia below the Sea of Marmara (Polat and Tuğrul, 1995) yields:
an intense pycnocline at ca. 100 m depth, although the depth of the

pycnocline has likely decreased in the past few decades (Murray et al., Bosporus NO3 input
   
1989). The surface 50 m has a salinity of about 17.9‰, which increases 11 3 −1 3 − −3
= 3:05 × 10 m yr × 7 × 10 cmol NO3 m
sharply to 21‰ at ca. 100 m and then more gradually to 22.5‰ below
15 − −1
1750 m. A suboxic layer of cold intermediate water (CIL) lies between = 2:1 × 10 cmol NO3 yr : ð12Þ
the oxic uppermost 50 m and the anoxic bottom water at 100 m
(Codispoti et al., 1991). The CIL forms largely in the northwestern The total import of dissolved NO−3 is therefore ca. 7.4 × 10
15
μmol yr− 1.
margins of the basin by winter cooling and spreads throughout the sea Although the two inputs are similar in magnitude, the enclosed
(Murray et al., 1991; Özsoy and Ünlüata, 1997). Oxygen and H2S go to environment of the Black Sea and shallow sill cannot be compared to
zero in this suboxic layer (Fig. 3). Nitrate, HPO2− and H4SiO4 reach open-ocean environments, where above we noted that the river input
4
maxima in the upper reaches of the layer (Fig. 7), NO− for an open-ocean environment was minor, even in coastal areas off
3 through
nitrification of settling organic matter before decreasing to zero via major rivers.
denitrification immediately above the anoxic bottom water (Codispoti This NO−3 input into the Black Sea is insignificant in comparison to

et al., 1991). As observed in the Cariaco Basin, anammox reactions the annual NO− 3 demand of phytoplankton. This demand can be

lower NH+ − estimated from the area of the sea of 420,000 km2 and the primary
4 and NO2 to zero at this same depth (Kuypers et al., 2003).
Bottom water in the basin is a mixture of Mediterranean water that productivity of 210 g Corg m− 2 yr− 1, assuming Redfield stoichiometry
enters through the Bosporus and the suboxic CIL, in the approximate for the organic matter (Redfield et al., 1963):
ratio of 1:3.3 (Murray et al., 1991). The penetration depth of much of −
Phytoplankton NO3 demand
the inflowing Mediterranean water is limited largely to the upper    
500 m, leaving the deeper basin waters progressively more isolated, ½ 11 2
= 4:2 × 10 m × 210 g Corg m
 
−2
yr
−1

with a RT possibly of several thousand years (Özsoy and Ünlüata, −


× 0:28 g NO3  0:36 g Corg
 
1997). The inflowing Bosporus plume delivers salt and O2 into the
deeper layers by the lateral injection of plume waters (Ivanov and
4 − −1
× 1:6 × 10 cmol NO3 g NO3


15 − −1
Samodurov, 2001) rather than by slow vertical diffusion. However, the = 1100 × 10 cmol NO3 yr : ð13Þ
flux of O2 is insufficient to prevent the built up of H2S, which has
concentrations currently reaching 100 μM at 200 m and increasing to The NO− −
3 demand is much larger than the external NO3 supply
380 μM at 2200 m depth. The bottom 300–400 m of the deep-water (Eqs. (11) and (12)) and can only be met by nutrient recycling in the
column, below 1750 m, is isothermal and isohaline (Murray, 1991). surface water and the vertical movement of waters from below the
The average export rate of waters into the Sea of Marmara through surface photic layer by upwelling and vertical mixing. This is
the Bosporus Strait, between the sea surface and 25 m, is ca. 605 km3 analogous to the nutrient supply to coastal waters of the open
yr− 1 (Özsoy and Ünlüata, 1997). Sub-surface water is imported into ocean, where upwelling oceanic waters represent the dominant
the Black Sea below this exported surface water to 40 m depth, at a source of nutrients, here equal to 110 × 1015 μmol NO−3 yr
−1
, or 10%
rate of 305 km3 yr− 1. An average fresh-water influx of 352 km3 yr− 1 is of the demand of primary productivity.
80 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

The rate of import of nutrients in support of new production is 3.2.3. Bottom-water residence time
simply the seawater flux necessary to deliver this amount of NO− 3 to The RT of the bottom water can be calculated from the oxidants
the photic zone, i.e., to the mixed layer lying immediately above the consumed in a parcel of water (Eqs. (15)–(17)) versus the rate that
CIL: Corg is oxidized, or the rate that Corg settles out of the photic zone. If
the small fraction of Corg that escapes oxidation in the water column
Upwelling ðmixingÞ rate and eventually accumulates on the seafloor is ignored, then
h  i
15 11 −2 −1
= 110 × 10  4:2 × 10 cmol m yr
−2
h i RT = ½ð635 − 490Þ + ð15 − 9Þ + ð1600Þmol Corg mðoxidants reducedÞ
3 −3 −1 ð14Þ
 8 × 10 cmol m = 33 m yr : −2 −1
½210 × 0:10  12mol m yrðrate of Corg oxidizedÞ
−2 −2 −1
This should represent a close estimate as the input from rivers and loss ¼1750 mol Corg m  1:75 mol Corg m yr = 1002 yr:
by the export of surface water into the Sea of Marmara may approxi- ð18Þ
mately balance. The RT of surface water is then ca. 1 to 2 years. As Lee
et al. (2002) reported a RT for water between 80 and 120 m in the Adjustments to the consumption of O2 (635 − 490) and NO− 3 (15 − 9)
suboxic layer of ca. 5 years, the 1 to 2 years seems a reasonable represent mixing of suboxic CIL with Mediterranean water in the ratio of
estimate for the photic zone. 3.3:1 (Murray et al.,1991) to form the bottom water. Unlike conditions in
the Cariaco Basin, both terms are small compared to the SO2− reduction
3.2.2. Oxidant balance in bottom waters term of 1600 mol Corg m− 2. The RT derived in this way is significantly
As we showed earlier for the Cariaco Basin, the consumption of the greater than the 387 years estimated by Murray et al. (1991), lies within
major oxidants during oxidation of Corg can be estimated from the the 1000 to 2000-yr range reported from radiocarbon measurements
utilization of dissolved O2, NO− 3 and SO4
2−
in the remineralization of (Ostlund, 1974), but is lower than the several thousand years advanced
organic matter as it settles to the seafloor. The O2 concentration in the by Özsoy and Ünlüata (1997), Grégoire and Stanev (2001), and Stanev
water imported from the Mediterranean is taken as 290 μM, its (2005) and the 2500 to 3000 years estimated by Anderson and Fleisher
approximate solubility at 8 °C and 35‰ (Weiss, 1970). The contribution (1991) from the U budget.
of NO− 3 reduction to the oxidation of organic matter is much less than An alternative estimate of the RT can be made from the export (or
that of O2 per unit volume of water, owing to the large difference in their import) of salt versus the salt inventory of the basin, i.e., the rate of salt
respective concentrations in the imported water. The maximum NO− 3 replacement, a condition for steady state.
concentration in the Black Sea (8 μM) is at about 60 m depth, the upper
part of the CIL, or suboxic zone. As noted above, the maximum is due to RT = ½volume × mean salt concentration  ½salt export rate
   
oxidative regeneration (nitrification) of settling organic detritus 5 3 3 −1
= 5:3 × 10 km × 22:2x  605km yr × 17:9x = 1100yr:
(Codispoti et al., 1991) and import of water from the Sea of Marmara,
below which it and the other forms of fixed-N go to zero at the suboxic– ð19Þ
anoxic interface. The SO2−4 contribution to the oxidation of Corg in the
bottom water is given by the concentration of H2S (380 μM). The The RT of the surface and near-surface waters will be less than this
sequence of reactions for the oxidation of Corg can be quantified in terms value, whereas the RT of bottom water will be greater. Indeed, Lee et al.
of the reduction of major oxidants in the lower 2150 m of the water (2002) reported that the RT of water in the upper part of the water
column, i.e., below the photic zone, as was outlined for the Cariaco Basin column increased to 625 years with depth through the uppermost
(Eqs. (3)–(5)). We again ignore the flux of H2S into the suboxic water 500 m. Although this and the other measurements fail to constrain
and its subsequent oxidation. tightly the RT of bottom water, the estimates summarized here suggest
that the mean RT lies in the range of 1500 to 2500 years.
Oxic respiration :
h i 3.2.4. Trace-metal budgets in bottom waters
−3 −1
0:290 mol O2 m × 2150 m × 1:0 mol Corg mol O2
The distributions of trace metals in the Black Sea water column are
−2
= 635 mol Corg m : ð15Þ similar to their distributions in the Cariaco Basin (Fig. 7). Dissolved
Mn and Fe display maxima just below the oxic–anoxic interface, due
Suboxic respiration : to the reductive dissolution of particulate oxyhydroxides (as Mn(IV)
   i and Fe(III)) settling from the oxic near-surface waters into the anoxic
½ 0:008 mol NO−
3 m
−3
× 0:70 × 2150 m × 1:25 mol Corg mol
−1 −
NO3 water (Brewer and Spencer, 1974; Haraldsson and Westerlund, 1991;
−2
= 15 mol Corg m : ð16Þ Lewis and Landing, 1992). Iron in the anoxic zone as Fe(II) is close to
equilibrium with FeS (mackinawite) and Fe3S4 (greigite) (Landing
Anoxic respiration : and Lewis, 1991), consistent with the removal of dissolved Fe as Fe
h   i “monosulfide”. Particulate Mn and Fe have complementary maxima
2− −3 −1 2−
0:38 mol SO4 m × 2100 m × 2 mol Corg mol SO4
at the base of the oxic zone, immediately above their dissolved
−2
= 1600 mol Corg m : ð17Þ concentration maxima (Brewer and Spencer, 1974). There may also be
direct supply of Fe to the central basin from the basin margins
Adjusting for units gives 1090 μM Corg as the sum of carbon added to (Wijsman et al., 2002; Lyons and Kashgarian, 2005). Dissolved Cu and
the bottom water. This value is only ca. 15% lower than the measured Zn decrease in concentration across the suboxic–anoxic interface
increase of CO2 in the bottom water over its concentration in the (Haraldsson and Westerlund, 1988, 1991; Lewis and Landing, 1992),
surface 100 m and 20% higher than that attributed to the oxidation of supporting the interpretation of their removal from seawater in the
Corg by Goyet et al. (1991). The excess can be attributed to entrainment anoxic zone, although they are undersaturated with respect to Cu2S
of CIL, the Black Sea water that mixes with imported Mediterranean and ZnS, respectively (Landing and Lewis, 1991). Cadmium and Pb
water in the ratio of 3.3:1 (Murray et al., 1991) to form the deep water also decrease in concentration across the suboxic–anoxic interface.
of the basin. The adjusted sum of oxidants reduced in the water Both are close to saturation with respect to CdS and PbS, respectively;
column is then equivalent to the CO2 generated from the oxidation of they are likely removed from solution as these, or similar, phases
Corg. Seventy percent of the Corg in the Black Sea is oxidized by SO2−
4 , (Landing and Lewis, 1991). Nickel exhibits little variation with depth
compared to 50% in the Cariaco Basin (Eq. (5)). (Brewer and Spencer, 1974; Haraldsson and Westerlund, 1988),
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 81

although a ca. 50% decrease is reported by Landing and Lewis (1991). an adsorbed phase (Jacobs et al., 1987; Landing and Lewis, 1991). Its
It has a uniform profile in other anoxic basins (Jacobs et al., 1985, export from the photic zone then equals its depositional rate as a
1987). We have no explanation for the difference in the measure- seawater-derived biogenous fraction (Piper, 1994), or 1050 μg Ni m− 2
ments of Ni in the Black Sea by Landing and Lewis versus yr− 1. At a mean bulk-sediment accumulation rate of 38.5 g m− 2 yr− 1
measurements by the other groups, although the weight of evidence (Calvert et al., 1991), this gives the following seawater contribution to
from this and the other basins favors the invariant vertical its sediment concentration:
concentration profile for this element in anoxic basins.
Dissolved Co shows an increase in concentration coincident with Marine Ni sediment concentration
the dissolved Mn and Fe maxima due to its release from dissolution of −2 −1 −2 −1
= 1050 cg Ni m yr  38:5 g m yr = 27 ppm: ð22Þ
oxyhydroxide phases, mostly of Mn, settling out of the oxic–suboxic
zone (Haraldsson and Westerlund, 1988, 1991; Lewis and Landing, Of the Ni in the surface sediment of 122 ppm (Calvert, 1990), 22% is
1992). In both the photic zone and anoxic zone, its concentration is derived from the photic zone and delivered to the sediment by a
low (ca. 0.2 and 0.25 nM, respectively, versus 1 to 4 nM between 200 biogenous vector. Measurements of the accumulation rate of Corg in
and 400 m). It is undersaturated with respect to simple sulfide phases. the surface sediments show that only about 1% of primary productiv-
Although some of the Co is associated with particulate Al2O3 in ity is buried (Calvert et al., 1991). This is reflected by the higher Corg:Ni
bottom-water suspended phases, some fraction is apparently ratio in plankton (0.020 g μg− 1) than in the sediment (0.0019 g μg− 1).
adsorbed onto particulate Fe (Lewis and Landing, 1992). Even so, its Molybdenum and V have depth profiles in the Black Sea (Emerson
extremely low concentration in the water column and slightly and Huested, 1991) similar to those of Cd, Cu, and Zn, indicating their
elevated concentration in the bottom water precludes measurable removal from the bottom water. The Mo concentration in the bottom
Co enrichment in the sediments above a lithogenous input, even if water is as little as 10% its concentration in the surface and near
totally removed from the bottom water. surface water. Similar to the Cariaco Basin, no marine Mo fraction is
These observations suggest that Fe, Cd and Pb are deposited from captured in sediment traps with settling particulate debris in the
the anoxic waters of the Black Sea by precipitation of their solid anoxic waters of the Black Sea, whereas the metal is strongly enriched
sulfides. Copper and Zn, which also have much lower concentrations in the bottom sediments above a lithogenous contribution. Again, its
in the anoxic zone, are likely deposited as sulfide complexes (e.g. MeS removal from the water column is apparently restricted to the benthic
(HS)−) and adsorbed onto solid phases, possibly onto FeS2, both being boundary (Emerson and Huested, 1991, Crusius et al., 1996). None-
undersaturated with respect to their respective sulfides (Landing and theless, its depletion throughout the anoxic region of the water
Lewis, 1991; Lewis and Landing, 1992). Finally, Co and possibly Ni are column suggests that the entire water column contributes to its ele-
undersaturated and probably exist as dissolved sulfide complexes in vated concentration in the sediment. Vanadium, as discussed earlier, is
the anoxic waters; their possible removed from the anoxic water is present as VO2(OH)2− 3 in oxic seawater and is reduced to the much
insignificant relative to their accumulation on the seafloor. That some, more strongly adsorbed VO(OH)− 3 under suboxic and anoxic condi-
but not all, redox-sensitive metals are removed by particle scavenging tions (Sadiq, 1988; Wehrli and Stumm, 1989; Emerson and Huested,
in anoxic waters, rather than by direct precipitation of their sulfide 1991), leading to its removal from the water column, its presence in
phases (Haraldsson and Westerlund, 1991), is supported by theore- sediment traps, and its deposition on the seafloor by adsorption reac-
tical studies of Dyrssen (1985). And finally, deposition of Ni from the tions under both suboxic and anoxic redox conditions.
water column is tied to the Corg cycle, i.e., to primary productivity. Although they are enriched in the sediments above a lithogenous
Similar to the calculations advanced for the Cariaco Basin, the input, the accumulation rates of the hydrogenous fractions of both
fluxes of several trace elements through the water column to the Mo and V from the bottom water are unexpectedly low, Mo more so
sediments as: (1) a biogenous fraction (Ni); (2) a hydrogenous than V. For Mo, its import rate from the Sea of Marmara, in which its
fraction added to the sediments through inorganic redox reactions (V concentration is 10 ppb (Pilipchuk and Volkov, 1974), is:
and Mo); or (3) by both mechanisms (Cd, Cu, Zn) are determined from
 
their supply and removal rates from the water column and an 3 −3
Mo import rate = 10 × 10 cg Mo m
understanding of redox processes within the water column and the  
3 −1 9 3 −3
bottom sediments. The rate that the micro-nutrient Ni is imported × 305 km yr × 10 m km
 
5 2 6 2 −2
into the photic zone (expressed as an upwelling rate) is simply:  4:2 × 10 km × 10 m km
= 7260 cg m
−2
yr
−1
: ð23Þ
−1 3 −3
Ni upwelling rate = 33 m yr × 0:49 × 10 cg Ni m
= 16; 200 cg m
−2
yr
−1
: ð20Þ Its concentration profile in the bottom water suggests that greater than
ca. 90% of its import is eventually deposited on the seafloor, correspond-
Loss of Ni from the photic zone via settling of Corg, namely the export ing to a removal rate from the anoxic water of 6500 μg m− 2 yr− 1. Its
of 10% of primary productivity, the global export rate as was adopted accumulation rate, however, is a small fraction of this amount.
earlier, or 21 g Corg m− 2 yr− 1 with the Ni:Corg concentration ratio of    
18 ppm:36%, is: −1 −2 −1
Mo accumulation rate = 50 cg g × 0:85 × 38:5 g m yr
  = 1640 cg Mo m
−2
yr
−1
; ð24Þ
−2 −1 −1
Ni loss = 21 g Corg m yr × ð18  0:36Þcg g
−2 −1
= 1050 cg Ni m yr : ð21Þ where the bulk Mo concentration in the surface sediment is ca. 50 ppm
(Volkov and Fomina, 1974). Approximately 10% is contributed by the
Again, Ni is not a limiting nutrient, although it is an essential nutrient. lithogenous fraction, based on the Mo:Al2O3 ratio of WSA and the Al2O3
For a bottom-water RT of 1500 to 2500 years, we might have concentration in the sediments, and 5% is contributed by biogenic debris,
expected to see as much as a doubling of Ni in the bottom water in based on a maximum Mo:Corg ratio of phytoplankton (Brumsack, 1986)
response to the remineralization of organic matter settling through and the rate of deposition of Corg obtained from the accumulation rate of
the water column. The lack of such a build up in this and other anoxic marine Ni.
basins (Jacobs et al., 1985, 1987; François, 1988; Haraldsson and The Mo accumulation rate then represents only 20 to 25% of the
Westerlund, 1988, 1991) suggests that the organically bound Ni input rate through the Bosporus. To account for the discrepancy, most
exported from the photic zone is deposited on the seafloor largely as of the Mo imported into the Black Sea must be exported back into the
82 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

Sea of Marmara in the surface outflow (Deuser, 1974). In support of we have elected to model the upper Jurassic KCF of northwestern
this interpretation, the maximum concentration of Mo in the water Europe (Figs. 1 and 2) in order to derive levels of primary production,
column of the Black Sea (ca. 110 nM) is in the very surface water water column redox, and hydrography in the Kimmeridge Sea during
(Emerson and Huested, 1991), the water that is exported back into the deposition of the cyclical organic-rich/organic-poor sediment
the Sea of Marmara under the estuarine circulation of the basin. As lithologies of this deposit. The KCF in England and correlative
little as 20% of the imported Mo from the Sea of Marmara must be formations in northern France are some of the most thoroughly
deposited on the seafloor (1450 μg m− 2 yr− 1) to achieve agree- examined black shales in the world, owing to the large oil source-rock
ment with its accumulation rate. Vanadium has a similar seawater potential of coeval rocks in the North Sea grabens. A comprehensive
profile to that of Mo (Emerson and Huested, 1991). As the chemistry review of this literature is beyond the scope of this paper. Rather,
of the uppermost part of the water column has been drastically following a brief summary of the principal features of the formation,
altered in the past several decades (Halim, 1991), it is not now pos- we present our interpretations of the water-column dynamics under
sible to know the natural concentration of trace elements in the which the KCF was deposited in the distal environment of the
oxic water in previous years, but the trace elements would also have Cleveland Basin in Yorkshire, England and then compare this
been exported largely back into the Sea of Marmara by the estuarine environment with the time correlative units deposited in the proximal
circulation. environment of the Boulonnais Basin in northern France. The data
The disparity between the inputs and accumulation rates of these have been published by Tribovillard et al. (2005), who has kindly
metals apparently reflects the small fraction of imported water from provided access to this valuable data set. One set of samples was taken
the Sea of Marmara that actually ventilates the Black Sea at depth. The from cores at three sites in the Cleveland Basin; the second set of
sill depth in the Bosporus must play a central role in limiting the net samples was collected from the correlative formations that outcrop in
flux of Mo, V and other trace metals into the bottom water such that the cliffs along the coastline immediately north of Boulogne. One
the import of these metals is only slightly greater than their export in centimeter-thick samples were taken at a spacing of 3 to 5 samples per
the surface water. 10 m of section in the Cleveland Basin and ca. 6 samples every 10 m at
the Boulonnais site. Two sections in the cores from the Cleveland
3.3. Summary Basin, referred to as primary cycles, were sampled on a centimeter
scale.
Agreement between the uptake of Ni in the photic zone by As we have discussed previously, possible bathymetric or paleo-
plankton and its deposition on the seafloor, in both the Cariaco Basin ceanographic differences between modern and ancient basins should
and Black Sea, supports its use as a proxy for the rate of primary not invalidate a comparison of their geochemical records. The
productivity in ancient basins whose bottom waters were O2 depleted. processes that are central to the relationship between the geochem-
Likewise, the estimates of the accumulation of Mo and V in the two istry and hydrography of the water column and the accumulation of
modern basins support their use as proxies of bottom-water redox. Corg and trace metals on the seafloor are quite separate from the
The 4 to 5-fold higher accumulation rates of Mo and V on the seafloor bathymetry of the basin, whether the “basin” had an actual
of the Cariaco Basin than in the Black Sea are probably due to bathymetric expression or was an epeiric sea, possibly with no sill.
differences in the circulation of the two basins that may largely reflect However, the texture and fine-scale laminar nature of the KCF, lack of
their sill depths. Both basins have an estuarine circulation. In the Black extensive bioturbation, and preservation of delicate articulated
Sea, the imported water enters at a depth of 25 to 40 m. Although bivalves commonly along bedding planes (Tyson et al., 1979) suggest
some fraction of this water clearly descends to depth, much of the a closer similarity of the depositional conditions of the KCF to the two
imported water must be mixed into the photic zone and exported back prominently-silled anoxic basins discussed here than to those of the
into the Sea of Marmara with its complement of trace elements. OMZs of modern continental margins.
Evidence for the mixing of the imported water and the CIL is the A comparison of the marine environment in the Cleveland Basin
maintenance of the salinity of surface water. The volumes of imported with the depositional environment in the Wessex Basin, approxi-
water from the Sea of Marmara and of riverine fresh water are mately 400 km to the south, where the type Kimmeridge section is
approximately equal, with salinities of 35.8‰ and 0‰, respectively; located (Cox and Gallois, 1981), is problematic. Whereas the
the salinity of a mixture of the two should be 17.9‰, which is the sedimentology and physical properties of the coastal outcrops in
salinity of the surface water (Özsoy and Ünlüata, 1997). An southern England have been studied extensively (e.g., Tyson et al.,
exceptionally low accumulation rate for both Mo and V on the 1979; Oschmann, 1988a,b; Wignall and Hallam, 1991; Wignall, 1994;
seafloor simply suggests loss of much of the trace-element inventory Tyson, 2004; Weedon et al., 2004), to our knowledge no analytical
of the imported water by the export of surface water, supporting the investigation has been carried out on these rocks comparable to this
suggestion that the mean bottom-water RT is likely significantly investigation of the Yorkshire cores. Likewise, outcrops in the
greater than 1000 years. Cleveland Basin are apparently of such quality that they allow for a
The contrast in sill depth between the Cariaco Basin (145 m) and much poorer examination of the physical properties of the formation
Black Sea (40 m) is likely critical to the mixing of imported waters. The than in the Dorset sections (Cope, 1974). A detailed examination of the
different hydrography of the two basins leads to the interpretation that chemical composition of drill cores from the Wessex Basin and the
ancient basins that exhibited O2 depletion in the bottom water for physical properties of cores from the Cleveland Basin will eventually
extended periods of time and that now have sediment concentrations of correct this discrepancy. In the meantime, we are aware that some of
marine V and Mo in the range of several 100 ppm probably had an our interpretations of the environment of deposition of the KCF, based
exchange of seawater with the open ocean possibly exceeding even that on the geochemistry of the cores from the Cleveland Basin, may not
of the Cariaco Basin, owing to the presence of a deep sill or no sill at all. apply to the conditions of deposition in the Wessex Basin, based on
the descriptions of outcrops and the paleontological record, more so
4. An ancient marine environment than on the geochemistry.

Piper and Isaacs (1994, 1996), Piper (2001), and Perkins et al. 4.1. General description
(2008) used the normative modeling approach described in this paper
to interpret the marine environments of deposition of black shales Deposits of the KCF in the Cleveland Basin represent one of the
ranging in ages from Late Devonian to Miocene, and Pleistocene most complete sections of this formation that accumulated in the
sediments from the Sea of Japan. Rather than reiterate these results, distal part of an epeiric sea during the Late Jurassic. It was but one
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 83

basin in an epeiric sea that extended from the Barents Sea to the and ferromanganese concretions are associated with rippled sand
English Channel (Hallam, 1975). Descriptions and interpretations of (Manheim et al., 1980). Seafloor erosion in the deep ocean has been
the stratigraphy, sedimentology, paleoecology, geochemistry and reported at many locations from sediment cores, seismic studies, and
paleoceanography of the formation can be found in Cox and Gallois bottom photographs (Johnson 1972; Kennett and Watkins, 1975;
(1981), Oschmann (1988a,b, 1990), Miller (1990), Wignall (1990, Ellwood and Ledbetter, 1979; Berger and Johnson, 1976). As an
1994), Macquaker and Gawthorpe (1993), Herbin et al. (1995), Proust example, relatively stiff Miocene sediment that hosts haloed burrows
et al. (1995), Tyson (1996, 2004), Wignall and Newton (2001), filled with Quaternary sediment is now exposed at the surface in the
Morgans-Bell et al. (2001), Tribovillard et al. (2001, 2005), Jenkyns central Pacific Ocean at greater than 5000 m depth. The burrows,
et al. (2002), Lees et al. (2004). The rocks are laminated at several which have a strong horizontal component, are truncated at the
horizons (Oschmann, 1988b; Wignall, 1991; Macquaker and seafloor (Piper et al., 1986, Fig. 4), requiring relatively recent sediment
Gawthorpe, 1993) and the preservation of delicate bivalves, 75% scour and presumably redeposition of the Miocene sediment, possibly
articulated, and restricted trace fossils commonly along bedding as rip-up clasts.
planes (Tyson et al., 1979; Wignall, 1994) suggest little sediment In the calculations that follow, we have assumed a depth of 300 m
reworking and tranquil bottom waters. The environment was for the Kimmeridge Sea; the uppermost 50 m of the water column
interrupted periodically by bottom scouring possibly from storm being oxic and the lower 250 m being O2 depleted, suboxic or anoxic
events, as evidenced for example, by rip-up clasts and hummocky for most of the time. A depth greater than roughly 300 m seems
cross stratification (Wignall, 1989). A dysaerobic benthic fauna unnecessary for the importation of nutrient-enriched seawater at
(Oschmann, 1988a; Wignall, 1990) is suggestive of a bottom water depth that supported primary productivity in the photic zone, as open-
strongly depleted in O2, similar to the water at intermediate depths on ocean seawater is enriched in nutrients at depths as shallow as 100–
the continental shelves of Peru (Codispoti, 1980), the Gulf of California 200 m (Fig. 4). A shallower depth than 150 m seems equally unlikely,
(Calvert, 1964), Namibia (Kuypers et al., 2005), and the Arabian because the benthic ecological evidence (Oschmann, 1988a,b; Wignall,
Peninsula (Codispoti, 1980), and to the seafloor in the Santa Barbara 1990) and the geochemical results presented here suggesting that
Basin within the California Borderland (Soutar and Crill, 1977). conditions of extreme O2 depletion in the bottom water were
Cycles of organic enrichment on several scales are a notable feature maintained for extended periods of time. This is not to suggest that
of the KCF. Five organic-carbon-rich intervals (ORIs) several meters the depth remained constant throughout the history of deposition at
thick are present in both basins, occurring at the same horizon, in the any one location, or that it was uniform throughout the basin at any
same ammonite zones (Herbin and Geyssant, 1993; Herbin et al., 1995; one time.
Weedon et al., 2004). Nested in both the ORIs and intervening Herbin et al. (1995) assigned a period of deposition for the ORIs of
organic-poor intervals (OPI) are units that Tribovillard et al. (2005) 500,000 years. They proposed that the 50 cm thick primary cycles
refer to as “primary cycles” up to 1.2 m thick. On an even finer scale are nested within the ORIs accumulated over a 25,000 year period. On this
millimeter thick laminae. As is the case with many other black shales, basis, the thicknesses of the three primary cycles sampled by
the high Corg contents of these rocks, the laminated structure, Tribovillard et al. (2005) and examined here of 60, 105, and 120 cm
restricted benthic fauna and occasional complete lack of benthic suggest a somewhat longer duration. Other analyses of sediment
fossils, together with the enrichment of several redox-sensitive trace cyclicity from the Wessex Basin suggest several different cycle lengths
elements above a lithogenous contribution have been interpreted as (Oschmann, 1988a,b). Weedon et al. (2004) reported a dominant cycle
indicating accumulation under O2-depleted bottom waters (Wignall of 38,000 years, which is close to the Milankovitch obliquity cycle of
and Myers, 1988; Tribovillard et al., 2005). Evidence for time intervals 41,000 years. This period is based more strongly on the measured
of high productivity, especially for some of the coccolith-bearing photoelectric factor than on Corg content as reported by Herbin et al.
limestone bands, has also been reported (Lees et al., 2004; Pearson (1995). House (1985) also reported a cyclicity of 41,000 years for the
et al., 2004; Weedon et al., 2004; van Dongen et al., 2006). upper part of the KCF. We take this as the cyclical duration of the
Deposition occurred in a series of basins (Hallam and Bradshaw, primary cycles in the Cleveland Basin.
1979), possibly shallow-silled, each with perhaps a different depth Assuming a complete sediment section in the Cleveland Basin,
that varied through time (Wignall and Hallam, 1991) in response to sediment thicknesses of the ORIs yield a compacted sediment
sedimentation, eustatic sea level changes, and basin subsidence and accumulation rate of ca. 0.0034 ± 0.0006 cm yr− 1. The primary cycles
emergence. It is not possible to establish the depth of the Cleveland had sediment accumulation rates ranging from 0.0015 (ORI 2-2) to
Basin with any certainty. Tyson et al. (1979) suggested a depth of a few 0.0026 cm yr− 1 (OPI 2-1) and the formation as a whole had an
hundred meters; Weedon et al. (2004) proposed a depth that varied accumulation rate of 0.0036 cm yr− 1 (Eq. (1)). The differences in
between 50 and 75 m; Wignall (1989), Stow et al. (2001), and Bjerrum accumulation rates between the primary cycles, the ORIs, and the
et al. (2001) proposed a depth less than 200 m, all based more or less entire KCF are likely due to a combination of variable dilution of the
on the depth of wave base and the presence of bottom-current and marine fractions by the lithogenous fraction and variable accumula-
storm-induced sedimentary features. tion of organic matter. For example, a single sample in primary cycle
Oschmann (1990) has questioned the prominence of high-energy ORI 2-2 has a Corg concentration of 31.4%, i.e., the highest accumulation
sediment features, owing to their possible over-representation rate of organic matter or lowest dilution factor, and lowest bulk-
relative to laminated sediment, whose record extending over sediment accumulation rate.
hundreds of years might be disrupted by a single storm event. A bulk rock density of 2.3 g cm− 3 (Weedon et al., 2004) yields mass
Additionally, wave base is perhaps an unreliable criterion on which to sediment accumulation rates of 35 and 60 g m− 2 yr− 1 for the two
establish the maximum depth of such basins, established on the primary cycles and 78 g m− 2 yr− 1 for the ORIs. The density of the
occurrence of sand ripples, rip-up clasts, and scour marks now sediment at the time of deposition was likely close to 0.3 g cm− 3, the
preserved in the KCF. Sediment structures indicative of bottom- approximate density of surface sediment in the Black Sea (Calvert et al.,
current activity in the KCF (Wignall, 1989) are observed at virtually all 1991) and the Cariaco Basin (Piper and Dean, 2002), giving sedimenta-
depths in the ocean today. Komar et al. (1972) discussed surface wave tion rates prior to compaction of 12, 20, and 26 cm per 103 years,
induced sediment transport to depths slightly exceeding 200 m depth. respectively. The estimates of mass accumulation rates allow the
Southard and Cacchoine (1972) demonstrated that breaking internal accumulation rates of Corg and trace elements and the dilution of the
waves at shelf-break and upper-slope depths can suspend sediment marine fraction by the lithogenous fraction to be determined, but only
that might then be transported by bottom currents laterally and to over an extended period of time. Even a single sample, as we noted
greater depths. Along the Blake Plateau at 775 m depth, phosphate earlier (Eq. (1)), encompasses as much as 280 years.
84 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

The Cap de la Crèche outcrop, located at the margin of the yr− 1 for the entire formation (Eq. (1)), yields the following estimate
Boulonnais Basin, comprises proximal facies (Fig. 2), including cross- for the accumulation rate of the lithogenous fraction alone:
bedded sand bodies (Herbin et al., 1995; Wignall et al., 1996;
Tribovillard et al., 2005), suggesting much shallower depths of Lithogenous faction accumulation rate
 
deposition than in the Cleveland Basin. The lithology of these units −2 −1
= 78 g m yr × ð9:4 ppm ThORI  16 ppm ThKCFmax: Þ
reflects a dynamic water column that may have resulted in the −2 −1
= 46 g m yr ; ð25Þ
removal of an indeterminate amount of sediment via seafloor
reworking and erosion, such that outcrops may retain an incomplete
representing an average of 60% of the bulk sediment. The maximum
sediment record. The enrichment of Zr relative to Al2O3 over the ratio
Th concentration in the KCF is 16 ppm (Fig. 5). This lithogenous input
for WSA (Fig. 8) further suggests the presence of coarse-grained
is roughly twice its input into the Black Sea (Calvert et al., 1991). The
sediment in the Boulonnais Basin, Zr being present as the heavy
source of this fraction to the Wessex Basin (Macquaker and
mineral zircon, which is concentrated in coarser-grained sediments
Gawthorpe, 1993) and possibly the Boulonnais Basin was the
due to the hydraulic equivalence between silt-sized zircon and sand-
London-Brabant Massif, whereas the source for the Cleveland Basin
sized quartz (Calvert and Pedersen, 2007). The sediments are
could have been the Grampian High (Fig. 1).
probably depleted in Corg, relative to primary productivity, and
The plots of Ni, Cu, V, Mo, and Zn versus Al2O3 demonstrate that
enriched in heavy minerals and quartz, compared with the model
curves drawn from the origin through the lowest metal:Al2O3 ratios
lithogenous composition.
have slopes that closely approach the metal:Al2O3 ratios of the WSA
The Argiles de Chatillon section (Fig. 2), which Wignall and
(Fig. 5). The curves are based largely, but not solely, on samples from
Newton (2001) characterize as a transgressive near shore facies
the Boulonnais Basin. As noted previously, Cu and Zn ratios with Al2O3
versus the maximum flooding facies of the Corg enriched units in the
are ca. 60% of the ratio for the WSA. However, their ratios closely
KCF (Tyson, 1996), is composed largely of dark marlstones and paper
approach those of the UCC standard, which contains 14.3 ppm Cu and
shales, with as much as 7.5% Corg (Tribovillard et al., 2001). Wignall
52 ppm Zn (Wedepohl, 1995).
and Newton (2001) conclude that it is perhaps the only section in the
The concentrations of the marine fraction of the trace elements, on
Boulonnais Basin that is environmentally correlative with the KCF.
a sample-by-sample basis, are represented by the displacement of
samples above the minimum curves. For Ni, its concentration in the
4.2. Lithogenous- and marine-sediment fractions: A trace-element marine fraction would have been less than 5% lower had the WSA
partitioning model value for the Ni:Al2O3 ratio been used to calculate marine Ni
concentrations instead of the slope of the minimum-value curve.
The composition of the lithogenous fraction of the Cleveland Basin The marine contributions of the other trace elements are similarly
KCF varies only slightly from the composition of crustal rock determined for each sample.
standards, just as the standards themselves vary to a minor degree
(Wedepohl et al., 1969–1978; Gromet et al., 1984; Taylor and 4.3. Biogenous sediment fraction: Primary productivity and rate of upwelling
McLennan, 1985; Condie, 1992). For example, the Al2O3 versus Th
relationship in the KCF, both elements being strongly partitioned into Primary productivity is determined from the accumulation rate of
the lithogenous fraction, exhibits a highly significant linear regression marine Ni, assuming 10% of primary productivity settles out of the
that extrapolates essentially to the origin and has a slope that closely photic zone (Suess, 1980; Pace et al., 1987; Berger et al., 1988; Thunell
approaches the Th:Al2O3 ratio in WSA (Fig. 5). We have no et al., 2000) and its Ni concentration is 18 ppm. Excluding the primary
explanation for the seemingly anomalously high Al2O3 concentrations cycle ORI 2-2, which is considered separately below, the average
in OPI 2-1 and no way of evaluating the offset from the other units, marine Ni concentration in the ORIs of 54 ppm (Table 1) and the
having no measurements of other elements that are partitioned average bulk-sediment accumulation rate translate into the following
equally strongly in the lithogenous fraction. However, the samples Ni accumulation rate:
from all other units define a single trend.
The ORIs, with a mean concentration of Th of 9.4 ppm (Table 1) and Marine Ni accumulation rate
bulk-sediment accumulation rate of 78 g m− 2 yr− 1, versus 83 g m− 2 −2 −1 −2 −1
= ½78 × 54cg m yr = 4200 cg Ni m yr : ð26Þ

The Ni accumulation rate equates to the following flux of Corg settling


out of the photic zone, for which the Corg concentration in organic
matter is 36% (Redfield et al., 1963):
h i
Corg settling flux = 4200 cg Ni m
−2
yr
−1 −
× 0:36 g g  18 cg g ½


−2 −1
= 84 g Corg m yr : ð27Þ

This yields a primary productivity of 840 g Corg m− 2 yr− 1, with 90% of


the nutrients recycled in the photic zone. This is similar to primary
productivity on the Peru Shelf (Chavez and Barber, 1987) and
somewhat higher than on the Cariaco shelf of 550 g Corg m− 2 yr− 1
(Goñi et al., 2003).
Miller (1990) considered that the productivity of the Kimmeridge
Sea was fairly low on the basis of an estimate of the long-term average
compacted Corg accumulation rate. Tyson (2004) obtained values for
the Wessex Basin of between 40 and 150 g Corg m− 2 yr− 1 and
Tribovillard et al. (2001) estimated primary productivity in the range
Fig. 8. The relationship between Al2O3 and Zr in samples from the Boulonnais Basin site.
of 52 to 175 g Corg m−2 yr−1 for the Argiles de Chatillon section in
Analyses of samples from the Yorkshire cores were not available. The caption to Fig. 6 the Boulonnais Basin. Using a high resolution cyclostratigraphic
lists the abbreviations of standards. estimate of the sedimentation rate of the KCF in the Wessex Basin,
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 85

Table 1
Summmary of the composition of sedimentary units examined in this study, from the KCF (Cleveland Basin in Yorksire, England), and age correlative units in Northern France
(Boulonnais Basin). All elemental concentration are in ppm; Corg is given in percent. Data were not available where none are listed.

Basin-interval Ammonite zone Th Cu Marine Ni Marine V Marine Mo Marine Corg Comments


(no. of samples) Cu Ni V Mo
Organic-rich intervals
Cleveland Basin-ORI 5 (6) Pectinatus/Hudlesstoni 8 48 38 86.3 59.7 89.4 6.5 25.0 26.9 8.2 Flixton core, depth 171 to 185 m
Cleveland Basin-ORI 4/3 (6) Hudlesstoni 8 48 38 88.0 62.6 118.7 45.6 32.6 31.6 9.5 Ebberston core, cycle 3,
depth 192.5 to 193.2 ma
Cleveland Basin-ORI 4 (6) Hudlesstoni/Wheatleyensis 10 54 40 81.8 46.4 135.7 44.0 18.0 16.4 9.1 Flixton core, depth 190 to 205 m
Cleveland Basin-ORI 3 (10) Elegans 10 49 35 66.5 30.7 124.4 32.4 11.9 10.2 6.7 Flixton core, depth 212 to 234 m
Cleveland Basin-ORI 2 (11) Autissiodorensis/Eudoxus 9 51 38 92.4 59.5 201.0 107.2 67.0 65.5 6.2 Flixton core, depth 252 to 268 m
Cleveland Basin-ORI 2/2 (16) Eudoxus 11 84 72 158.4 127.3 368.2 278.3 119.3 117.5 13.9 Marton core, cycle 2,
depth 260 to 260.6 ma
Cleveland Basin-ORI 1 (6) Eudoxus 10 66 50 112.0 72.4 166.8 53.9 71.0 68.3 7.1 Flixton core, depth 274 to 290 m

Organic-poor intervals
Cleveland Basin-OPI 5 (0) Hudlestoni Flixton core, depth 185 to 190 m
Cleveland Basin-OPI 4 (1) Scitulus 14 38 19 58.0 7.7 164.0 20.3 3.0 1.0 2.3 Flixton core, depth 205 to 212 m
Cleveland Basin-OPI 3 (3) Autissiodorensis 11 41 26 65.7 25.2 134.7 18.2 6.7 5.4 3.7 Flixton core, depth 234 to 252 m
Cleveland Basin-OPI 2 (0) Eudoxus Flixton core, depth 268 to 274 m
Cleveland Basin-OPI 2/1 (23) Eudoxus 10 71 50 104.2 50.7 165.5 23.3 5.6 4.8 5.7 Marton core, cycle 1,
depth 270 to 270.85a
Cleveland Basin-OPI 1 (5) Eudoxus/Mutabilis 13 41 24 65.2 14.7 164.2 41.0 12.0 10.3 (2.5)b 3.1 Flixton core, depth 290 to 350 m

Organic-rich, organic-poor intervals undifferentiated


Boulonnais Basin-(15) Elegans to Eudoxus 9 20 11 45 19 81 2.8 2.3 1 Argiles de Chatillon
Boulonnais Basin-(11) Pallassioides 7 12 4 28.8 5.9 73.4 4.2 0.5 1 Argiles de Wimereux
Boulonnais Basin-(21) Pallassioides to Eudoxus 8 14 4 39.1 11.2 78.5 3.4 1.5 1 1.9 Cap de la Creche, 102 m outcrop
a
Depths listed for samples collected from the Marton and Ebberston cores (primary cycles 1, 2, and 3) have been scaled to approximate depths within the Flixton core, as reported
in Tribovillard et al. (2005).
b
The marine Mo value of 10.3 ppm is skewed by a single value of 45 ppm; the mean otherwise is 2.5 ppm.

Weedon et al. (2004) obtained a long-term mean primary productiv- the bulk-sediment accumulation rate and the Corg concentration
ity for the Kimmeridge Sea of ca. 200 g Corg m− 2 yr− 1, with a range (mean = 8.7 wt.%).
from b50 to 840 g Corg m− 2 yr− 1. The highest values of ca. 700 to 840 g
Corg m− 2 yr− 1 were from the same ammonite zones as the ORIs in the −2 −1 −2 −1
Corg accumulation rate = ð78 × 0:087Þg m yr = 6:8 g m yr ;
Cleveland Basin, consistent with our estimate of primary productivity
for the ORIs based on a completely different approach. ð28Þ
The estimates by Tribovillard et al. (2001) and Tyson (2004) were
based on the calculated burial flux of organic carbon from the linear This yields a burial efficiency of 0.81% of primary productivity. The
sedimentation rate and the pre-compacted bulk density of the rock burial efficiency of Corg estimated by Tribovillard et al. (2001) was 5 to
and the losses of carbon during recycling and settling through the 8%. It has been commonly assumed that such high burial efficiencies
water column and on the seafloor. They are significantly lower than occurred under anoxic bottom waters, but the near total lack of
the rates of primary production on modern continental shelves accumulation of marine V and Mo in the Argiles de Chatillon section
(Walsh, 1981, 1991; Müller-Karger et al., 2005). Rather, they are suggests that the bottom water was oxic (see Section 4.4). The burial
characteristic of primary productivity in the central gyres of the open efficiency estimated from the level of primary productivity calculated
ocean (Berger et al., 1988; Behrenfeld and Falkowski, 1997). Under by Tyson (2004) for the Wessex Basin likely would have been higher.
such low levels of primary productivity, the photic zone extends to as The variable composition and bulk-sediment accumulation rates of
much as 200 m depth. A seafloor depth less 200 m in the Kimmeridge the rocks, together with uncertainties introduced by our assumptions
Sea would have placed the entire water column of the Wessex and of single (model) compositions for the lithogenous fraction and
Boulonnais Basins in the photic zone. Primary productivity would organic matter, again permit calculations only of mean rates of
have extended to the seafloor, offering the opportunity for a highly primary productivity for the ORIs. Calculations based on the
efficient utilization of nutrients, as well as high benthic production composition of individual samples are further limited by the sampling
(Jahnke et al., 2000, 2008). It seems unlikely that these low levels of scheme; namely, 1 cm samples taken at ca. 1 to 3 m intervals
primary productivity would have been maintained under such (Tribovillard et al., 2005), versus cyclical changes in the geochemistry
conditions. Focusing of sedimented organic matter (Tyson 1996) of the formation commonly on a decimeter scale (Tyson, 1996). This
would make our calculation of primary productivity a maximum has been overcome, but still only partially, by the sampling of two
value. However, it is unlikely that the high level of primary primary cycles on a 1 to 2 cm interval (see Section 4.5).
productivity estimated for the Cleveland Basin can be attributed The settling flux of organic carbon in the Cariaco Basin, collected by
totally to focusing, as the accumulation of marine Ni yields the same sediment traps moored 145 m above the basin floor at 1400 m depth,
high level of primary productivity for the ORIs in the Cleveland Basin is 1 to 2% of primary production (Thunell et al., 2000; Goñi et al.,
as was obtained by Weedon et al. (2004) for the ORIs in the Wessex 2003), similar to values reported for the oxic environment of the open
Basin. ocean (Pace et al., 1987). One half of this settling flux is buried; the
Further insight into the production, settling, and burial fluxes of bulk of the other one half is oxidized at the seafloor, yielding a similar
organic matter in the Cleveland Basin can be obtained from an burial efficiency to that estimated for the KCF. Canfield (1989)
estimate of the accumulation rate of Corg in the ORIs, the product of calculated a similar burial efficiency for the Baltic Sea, at a depth of
86 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

230 m and sediment accumulation rate of 0.023 cm yr− 1. The


sedimentation rate for the ORIs adjusted for compaction was 0.024 cm
yr− 1.
The burial efficiency of Corg in the Cariaco Basin, Black Sea, and
Baltic Sea, and at additional sites in oxic environments in the modern
ocean (Henrichs and Reebugh, 1987) can be summarized by the
compilation of carbon accumulation rates and the linear sedimenta-
tion rate (Fig. 9) for a wide range of depths and redox conditions
(Calvert et al., 1991). The data from the KCF lie close to the curve
presented in this figure, revealing that Corg preservation was not
enhanced in the Kimmeridge shales compared with its preservation in
modern marine sediments. The use of logarithmic axes in Fig. 9
naturally obscures a good deal of scatter in the data, but is useful as a
representation of several orders of magnitude with fixed relative
precision (Dehaene et al., 2008).
It is likely that local rivers delivered a small fraction of the nutrient Fig. 10. Relationship between marine Ni versus Corg. All samples are enriched in Ni,
demand of phytoplankton in the Kimmeridge Sea. An estimate of the relative to its concentration in marine plankton, shown by an asterisk (Collier and
area extending some 3000 km from the Barents Sea to the English Edmond, 1983). Sample symbols are listed in Fig. 5.
Channel would have been greater than 5 × 105 km2 (Oschmann, 1990).
If the mean level of primary productivity for the entire sea was only
half the level estimated from the Yorkshire core, primary productivity
would have been 2 × 1014 g Corg yr− 1, on average. The NO− 3 demand The Cumarine:Nimarine ratio is 0.56 (7.5 ± 1.4 ppm:13.4 ± 2.4 ppm),
would have been 2 × 1018 μmol NO− 3 yr
−1
using the mean C/N ratio of which is similar to the average ratio for plankton of 11:18, or 0.61
marine phytoplankton in Section 3.2.1. This demand would have (Collier and Edmond, 1983). The similarity of the ratios identifies a
required a fluvial input 380 times the fluvial input to the Black Sea single marine biogenous source for both elements and probable lack
(Eq. (11)), or more than 20 Amazon Rivers at its current NO− 3 of anoxic conditions in the Boulonnais Basin (see Section 4.4), which
concentration and discharge rate (Maybeck, 1982). would have led to an enrichment of Cu relative to Ni due to the
The relationship between Ni and Corg lends support to our additional trapping of Cu from anoxic bottom waters where Ni
contention that Ni can be used as a proxy for Corg deposition. The removal does not occur (Piper and Dean, 2002). The similarity of the
mean Corg:Nimarine ratio in the ORIs is 0.0019 g μg− 1 (Fig. 10). Although Cumarine:Nimarine ratio in the Argiles de Chatillon section to that of
exhibiting considerable scatter, it is similar to the ratios in sediment modern plankton further supports the model compositions of both
from both the Black Sea and Cariaco Basin of 0.0012 and 0.0013 g μg− 1, Jurassic plankton and the lithogenous fraction.
respectively. The higher ratio for plankton of 0.020 μg g− 1 (Collier and Importation of open-ocean seawater, necessary to support primary
Edmond, 1983; Eq. (27)) merely reflects the loss of Corg via its productivity, may have been restricted by the presence of islands and
oxidation in the water column and on the seafloor during burial shallow shelves in, and marginal to, the Kimmeridge Sea (Ziegler,
(Eq. (28)). 1990). However, the presence of a complex seaway between the
Added support for this method of estimating organic matter Arctic, Central Atlantic and Tethyan Oceans (Oschmann, 1990;
accumulation is provided by the relationship between Cu and Ni in the Bjerrum et al., 2001) would have allowed for free exchange of
marine fraction of sediment from the Argiles de Chatillon section, seawater between the Late Jurassic epeiric sea and the Arctic Ocean
which, as noted earlier, is perhaps the only sediment section in the and/or the newly opened Atlantic Ocean. It is possible that the islands
Boulonnais Basin that can be compared with the KCF in England and shoals actually enhanced primary productivity in the Kimmeridge
(Wignall and Newton, 2001). The Cumarine versus Nimarine trend of the Sea by disrupting the flow field of surface currents and enhancing
two trace nutrients at this site (Fig. 11) exhibits a regression with shallow upwelling in a manner similar to the enhancement of
r2 = 0.57, which increases to 0.77 with the omission of a single sample. production in the lee of submarine banks and oceanic islands today,

Fig. 11. Marine Cu versus Ni in samples from the Boulonnais Basin. The slope of the
curve represents the mean ratio in marine plankton (Collier and Edmond, 1983).
Fig. 9. The relationship between the accumulation rate of Corg and bulk sedimentation Concentrations less than 5 ppm should be disregarded, owing to the uncertainty
rates of modern basins. The curve is based on sediment samples from the open ocean imposed from modeling the elemental composition of the lithogenic versus marine
(Calvert et al., 1991). fractions (Fig. 5).
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 87

the so-called Island Mass Effect (Doty and Oguri, 1956; Heywood et al., shallowness of these epeiric seas would have facilitated mixing of this
1990; Perissinotto et al., 1992). high nutrient bottom water into the photic zone, such that many sites
The vast extent of high levels of primary productivity is similar to of elevated primary productivity would have been located in epeiric
its distribution in the ocean today (Behrenfeld and Falkowski, 1997). seas and not along the ocean margins, as is their location today.
In the northern Indian Ocean, primary productivity seasonally extends
oceanward from the upwelling center along the Arabian Peninsula
several 100 km south and east towards the center of the Arabian Sea 4.4. Hydrogenous sediment fraction: water-column redox and residence
(Brock and McClain, 1992; Ryabchenko et al., 1998; Dickson et al., time
2001; Wiggert et al., 2005). In the Atlantic Ocean, it extends an even
greater distance from centers of coastal upwelling along northwest We have shown that dissolved V and Mo are removed from the
(Huntsman and Barber, 1977; Barber and Smith, 1981) and southwest anoxic bottom waters and deposited in the sediments of the Cariaco
Africa (Nelson and Hutchings, 1983; Pitcher et al., 1992) and off Peru in Basin and the Black Sea, with V removal at a higher redox potential
the Pacific Ocean (Brink et al., 1981). In the North Atlantic elevated than Mo. This difference allows the V versus Mo enrichment in
levels of primary productivity extend from east and south of Iceland sediments above their lithogenous contributions to identify past
into the Norwegian Sea and North Sea, being highest in coastal areas conditions of suboxia (V) and anoxia (V and Mo) in the KCF
of Ireland and Great Britain (Taylor et al., 1993; Stenseth et al., 2004). It sedimentary sequences. The much higher marine V:Mo ratios versus
is seasonally elevated also throughout several inland seas, e.g., the the concentration of marine Mo in the Cleveland Basin section
Gulf of California (Hidalgo Gonzalez and Alvarez Borrego, 2004; (Fig. 12), compared with their values in the sediments in the Black Sea
Douglas et al., 2007), the East China Sea (Gong et al., 2003), Yellow Sea and Cariaco Basin that are accumulating under fully anoxic conditions,
(Behrenfeld and Falkowski, 1997), and the Sea of Okhotsk (Sorokin, suggest that the bottom waters in the Cleveland Basin varied from
1999). It is similarly high in the Baltic Sea, but anthropogenic nutrient suboxic to anoxic during deposition of the ORIs, possibly achieving
inputs into this and the other ocean margin seas limit this comparison. oxic conditions for brief periods of time, and from oxic to suboxic
The elevated levels of primary productivity that are characteristic of during deposition of the OPIs. For most of the depositional history,
these and other ocean margins require the importation of nutrients bottom-water conditions were suboxic. The high concentrations and
into the photic zone from depths of a few tens to ca. 200 m. accumulation rates of marine Mo (100 to 200 ppm) and marine V (200
The elevated sea level in the Late Jurassic that flooded vast areas of to 300 ppm) in many samples from the ORIs further suggest free
northern and western Europe (Ziegler, 1990; Blakey, 2001; Scotese, exchange of bottom water with the open ocean. Unlike the major
2001) probably contributed to the importation of nutrient-enriched
seawater at depth from the open ocean. Haq (1987) has estimated a
sea level rise of 150 m. Studies of the Russian Platform (Sahagian
et al., 1996) place the rise at 50 m, although more recent work has
questioned the way in which past sea levels have been determined
(Moucha et al., 2008). Regardless, the increase in sea level likely
established a connection between the Kimmeridge Sea and the global
ocean. A connection to the open ocean only 150–200 m deep would
have routed major ocean-boundary currents onto the continent.
These currents today—the Gulf Stream, Canary Current, Benguela
Current, Agulhus Current, Peru Current, California Current, Kuroshio
Current—are restricted to the margins of the ocean basins (Sverdrup
et al., 1942). Together with the equatorial upwelling systems, they
account for 75% of global ocean primary productivity (Chavez and
Toggweiler, 1995). Their incursion onto the continents in Late Jurassic
time would have transferred into epeiric seas the enormous
inventory of nutrients they transport between ca. 100 and 300 m
depth. This importation of nutrients would have supported high
primary productivity at sites in the Kimmeridge Sea, West Siberian
Basin, and other epeiric seas. The sea-level rise of only 50 m would
also have enhanced trapping of terrigenous sediment in river
estuaries, limiting dilution by the accumulating lithogenous material
(Eq. (25)). Sea level reached a maximum during deposition of ORI-4,
Fig. 12. The relationship between Mo and V:Mo ratios in the marine sediment fraction of
but a high-stand may have been maintained throughout deposition the KCF. See Fig. 2 for the location of the OPIs and ORIs in the KCF. See Fig. 5 for the
of all of the ORIs (Miller et al., 2005), owing to seafloor subsidence in identification of sample symbols. Colors define the different bottom-water redox
response to crustal rifting (Ziegler, 1990; Graversen, 2002), as well as conditions as labeled. For the purpose of constructing this diagram, i.e., to show all values
eustatic sea-level rise. as positive, marine Mo and V concentrations that are less than 2 times the detection limit
(1 ppm) are assumed to be 1 ppm. All marine V:Mo ratios less than 0.5 are set at 0.5.
The importance of upwelling zones for the deposition of petroleum Samples with Mo concentrations less than approximately 10 ppm have a V:Mo ratio
source beds, an idea that dates back to Trask (1932), was examined by between 0.5 and 100. Most of these samples are from the OPIs. Such ratios suggest
Parrish (1987) using global paleoclimate models and the distribution deposition under oxic to suboxic conditions. An oxic area that includes these samples is
of rock lithologies. Although the boundary conditions used by Parrish included for illustrative purposes only, as hydrogenous V and Mo concentrations would be
zero in sediment deposited under oxic redox conditions, although the two should have a
have been questioned by Tyson (1995), Parrish found statistically
weak biogenic input. Samples with a marine Mo fraction greater than 15 ppm define the
significant numbers of co-occurrences of organic-rich facies and anoxic region and the suboxic-to-anoxic region. A suite of samples from a single outcrop
predicted sites of upwelling zones for much of the Phanerozoic but the typically plots along a single concave-up curve defined by decreasing V:Mo ratios with
correspondence was not uniform. Notably, there was no link between increasing Mo concentrations. This distribution is seen for the Permian Phosphoria
the formation of organic-rich facies and proposed upwelling zones for Formation (Piper et al., 2007) and the Devonian–Mississippian Ohio-Sunbury Shale of the
Appalachian Basin (Perkins et al., 2008). The trend identifies samples deposited under
the Late Jurassic. One possible explanation for the failure of a anoxic conditions as having V:Mo ratios approaching that of seawater, whereas samples
correlation may have been the transfer of major nutrient reservoirs deposited under suboxic conditions have elevated V:Mo ratios, owing to the different redox
onto the continents, as proposed here, owing to the sea-level rise. The conditions under which the two accumulate (Emerson and Huested, 1991).
88 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

nutrients, which are recycled within the photic zone, the accumula- under anoxic conditions (Jacobs et al., 1985, 1987) would have
tion of these redox-sensitive trace elements from bottom water and elevated the Cu:Ni ratio in the marine-sediment fraction, relative to
the trace nutrient Ni from the photic zone would have required the ratio in plankton.
continuous renewal of bottom waters to support their flux to the Sulfidic conditions within the shallow sediment horizons and
seafloor. Any barrier that might have restricted advection of water into extending into the bottom water likely created ideal conditions for
the Kimmeridge Sea was likely no shallower than the sill to the Cariaco organic matter preservation via sulfurization (Sinninghe Damsté and
Basin. If shallower, their concentrations and accumulation rates would de Leeuw, 1990; Lallier-Vergès et al., 1997; van Dongen et al., 2006)
have approached values now observed in the Black Sea. There may and the hydrogenation of biomolecules by H2S (Hebting et al., 2006).
have been no sill at all, similar to the bathymetry of the Gulf of This draws attention to the care that is needed when using
California, as the exchange of bottom water with the open ocean was sedimentary proxies of anoxia to distinguish between the chemical
more dynamic (Eq. (29)) than in either the Black Sea or Cariaco Basin, state in the pore water of accumulating sediments and that of the
and primary productivity (Eq. (27)) was significantly higher. The water column. In relatively shallow-water marginal marine settings
prevalent suboxic–anoxic condition of the Kimmeridge bottom waters today, the sediment pore water is commonly anoxic, whereas the
with occasional oxic periods is consistent with the dysaerobic benthic bottom water is only occasionally anoxic, likely similar to the
faunal evidence of Oschmann (1988a,b) and Wignall (1990), as well as condition in the Kimmeridge Sea, where the bottom water was
the range of the Indicator of Anoxicity developed by Raiswell et al. dominantly suboxic (or dysaerobic; Wignall, 1991).
(2001), whose values are consistent with the assignments of the
oxygen-restricted biofacies identified by Wignall and Hallam (1991). 4.5. Hydrography of primary cycles
As the lower part of the water column in the Cleveland Basin was
largely suboxic to anoxic throughout deposition of the ORIs, the The detailed sampling of two primary cycles, one in OPI 2 (OPI 2-1)
bottom-water RT should be given by the inventory of V in the water and one in ORI 2 (ORI 2-2), offers the opportunity to compare the
column and its rate of accumulation on the seafloor. If we assume a V hydrography of the water columns of these two apparently different
concentration in seawater of 1.8 ppb (Collier, 1984) and its removal environments of deposition (Fig. 13). A relatively uniform rate of
from strongly O2 depleted bottom water of about 10% (Emerson and flushing of bottom water during deposition of the primary cycle ORI
Huested, 1991; Piper and Dean, 2002), the average V concentration in (2-2) is suggested by the concentration of marine V. In the 16 samples
the marine fraction of the ORIs of 80 ppm gives the following mean RT: that were collected from this cycle, marine V averages 270 ppm
(Tribovillard et al., 2005) and varies over the rather narrow range of
h i 195 to 320 ppm. Bottom-water RT varied from 4.0 to 6.6 years
3 −2 −1
RT = 250 × 10 L m × 1:8 cg V L × 0:075
(Eq. (29)), assuming a linear rate of sediment accumulation. Marine
h i
−2 −1 −1 Mo varies between 25 and 220 ppm and the V:Mo ratio varies from
 78 g m yr × 80 cg g = 5:4 yr: ð29Þ
0.84 to 54, with a single high value at the very top of the section
(Fig. 13). Omitting this one value, the ratio averages 3.3 ± 2.7. The
The efficiency of V removal from the bottom water of the trend of the ratios, lower in the center of the unit than at the top and
Kimmeridge Sea under the suboxic to anoxic conditions is, of course, bottom, are indicative of a gradual change in bottom-water redox. The
not constrained, but for the purposes of this analysis we have taken higher ratios at the top and bottom are indicative of bottom water
what may be a conservative value of ca. one fifth its removal from the suboxia at these times (V:Mo ratios greater than ca. 3), whereas the
bottom waters in the Cariaco Basin (Emerson and Huested, 1991). bottom water was strongly anoxic throughout deposition of the
Likewise, the concentration of V in Jurassic seawater could have varied middle part of the section. The high Corg and Ni concentrations in the
from its concentration in the ocean today. Finally, selecting a seafloor middle of the section suggest that primary productivity was a major
depth of 200 m rather than 300 m reduces the RT to 3.2 years, which factor in varying the redox conditions.
still would have precluded seasonal overturns. Calculations of primary productivity (Eqs. (26) and (27)), based on
Wignall (1989, 1994) presents evidence of flushing events in the the concentration of marine Ni of 120 ± 55 ppm, averaged 840 ± 380 g
Kimmerdige Sea by oxic seawater, possibly lasting several years Corg m− 2 yr− 1. The upper limit exceeds measured values in the oceans
separated by periods during which the bottom water was suboxic to today by ca. 45%. A 2.7-fold increase in Th from the center of the cycle
anoxic. Although flushing events more or less continuously on the to the margins suggests that the high Ni concentration in the central
scale of seasons seems unlikely, the low mean sediment accumulation part of the cycle may reflect in part lower dilution of the marine
rate of 0.0034 cm yr− 1 and bottom-water RT of a few years for the fraction by the lithogenous fraction than at the top and bottom of the
ORIs can account for the presence of epifauna coexisting with textural cycle. Marine V also exhibits slightly lower concentrations in this
and geochemical signals of extreme O2-depleted, even anoxic, interval.
conditions. Bottom-water flushing events likely were more prevalent The chemical composition of a primary cycle within OPI 2 (OPI 2-1)
during deposition of the OPIs, leading to oxic bottom waters that may was also determined on a centimeter scale (Tribovillard et al., 2005).
have persisted over centuries. None of these rules out the possibility of The concentration of marine V is less that 5 ppm in 12 of the 23
storm conditions that might have briefly interrupted suboxic to anoxic samples (Fig. 13). The marine Mo concentration varies from less than
conditions, even in the ORIs. 1 ppm to 35 ppm, but is less than 3 ppm in 8 of the same samples
The marine V and Mo values of samples from the Argiles de (Fig. 13). The concentrations of both metals in these samples are
Chatillon site in the Boulonnais Basin plot in the oxic to oxic-suboxic simply below the limit of the modeling scheme. Within the center of
field (Fig. 12); marine V and Mo concentrations are essentially zero in the cycle, the marine V:Mo ratios are elevated, owing to V en-
most samples (Table 1). Although their low concentrations might be richment. This trend suggests that the bottom water varied from oxic
explained by a limited volume of suboxic to anoxic water in this toward the top and bottom of the cycle (low concentrations of both
shallow marginal setting (Wignall and Newton, 2001), the near total elements) to suboxic in the center (elevated marine V:Mo ratios).
absence of Mo, which can accumulate weakly from seawater on open Water column chemistry was significantly less stable than during
continental shelves under suboxic conditions (McManus et al., 2006), deposition of ORI 2-2. Similar to ORI 2-2, marine Ni and Corg are
supports dominantly oxic bottom-water conditions in the Boulonnais elevated in the center of the cycle. The average marine Ni
Basin. The similarity of the Cu:Ni ratio in the marine fraction (Fig. 11) concentration of 40 ± 37 ppm corresponds to a primary productivity
to the ratio in plankton (Collier and Edmond, 1983) also supports oxic of 480 ± 440 g m− 2 yr− 1, significantly lower than primary produc-
to suboxic bottom-water conditions, where the accumulation of Cu tivity of the ORIs. The results of the analyses of these two primary
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 89

Fig. 13. Relationship between Corg, marine Ni concentrations, and the marine V:Mo ratio in two primary cycles, one from ORI 2 and one from OPI 2. A third primary cycle (ORI 4-3) has
a thickness greater than 120 cm. It was sampled only sparingly (Table 1), but trends in its geochemistry parallel those in ORI 2-2. The main difference between the two is lower
concentrations of Corg, Mo, Ni, and V in ORI 4-3 than in ORI 2-2; in both primary cycles, half of the samples have a V:Mo ratio less than 2. The individual cycles are interpreted as
accumulating over 41,000 years (see text). In OPI 2-1, the filled diamonds represent samples in which the marine V concentration is less than 5 ppm and the marine Mo concentration
is less than 2 ppm; their concentrations are less than the modeling scheme is able to calculate. These samples likely accumulated under oxic conditions. Alternatively, their low
marine concentrations may reflect dilution by the lithogenic fraction. The filled squares represent samples for which marine Ni is less than 5 ppm. Note the two x-axes. Greater scatter
to the data for the OPI 2-1 than for the ORI 2-2 suggests less stable water column conditions in the former. In both units, but more so for the primary cycle ORI 2-2 (and ORI 4-3), both
Ni and Corg track V:Mo ratios, suggesting that primary productivity was the major factor that controlled the concentration of Corg and bottom-water redox, in the former case more so
than dilution by the lithogenous fraction and in the latter case more so than advection.

cycles support our interpretations of the ORIs that were sampled at Accumulations of redox-sensitive trace elements (V and Mo) in the
much less frequent intervals. ORIs recorded O2 depleted conditions that, except for relatively brief
periods when the water column was oxic throughout, extended from
5. Summary and conclusions the sediment pore water into the water column itself for extended
periods of time. Although redox conditions varied from oxic to anoxic,
The chemistry and hydrography of the water columns and the suboxic conditions were the most prevalent, consistent with the
geochemistry of the Holocene sediments of two modern anoxic basins— widespread occurrence of dysaerobic biofacies throughout much of
the Cariaco Basin and the Black Sea—have been used to model the the KCF and the rarity of horizons entirely lacking benthic fossils
processes responsible for the deposition of a group of redox-sensitive (Oschmann, 1988a; Wignall, 1990). Similar to the nutrient demand in
trace metals and trace nutrients in the two basins. By making several the photic zone, the accumulation of these trace elements required
simplifying assumptions, the model has then been applied to hindcast renewal of bottom water on the scale of a few years. The suboxic state
the hydrography and geochemistry that prevailed during the deposition of the Kimmeridge Sea bottom waters should not be considered rare
of the Late Jurassic Kimmeridge Clay Formation in the Cleveland Basin, or unusual. Denitrifying conditions occur over vast areas of modern
Yorkshire, England. The results reveal that primary productivity in the continental margins, e.g., beneath areas of high primary productivity
basin during deposition of five ORIs was ca. 840 g Corg m− 2 yr− 1, the in the Arabian Sea (Ganeshram et al., 2000), the continental slopes of
same as its rate on the Peru shelf today. The deposit currently has the Gulf of California (Calvert, 1964), western Mexico (Codispoti,
exceptionally high concentrations of Corg, although less than one percent 1980), the California Borderland basins (Soutar and Crill, 1977), and
of primary productivity escaped oxidation during deposition, burial, and continental shelves off Namibia (Bailey, 1991) and Peru (Arthur et al.,
diagenetic losses. Evocation of high planktonic productivity as the first- 1998). Combinations of high organic carbon fluxes and restricted
order control of the organic enrichment of the KCF is in accordance with ventilation cause these environments to be poised at this suboxic
the conclusions of many workers (Huc et al., 1992; Bertrand et al., 1994; state. Under this state both NO− 3 (40 μM) and O2 (50 μM) act as
Ramanampisoa and Disnar, 1994; Tribovillard et al., 1994; Saelen et al., oxidants. There seems no reason why a similarly vast expanse of this
2000; Jenkyns et al., 2002; Tribovillard et al., 2005). hydrographic condition should not have occurred in the ocean at
To sustain the high level of primary productivity, the basin relatively shallow depths in the past.
required the importation of nutrient-enriched seawater from the Calculations of the nutrient and trace-metal budgets for the KCF
open ocean at a depth below the photic zone no greater than 50– emphasize the dominant role of the lateral supply of nutrients required
250 m. Nutrients were delivered to the photic zone by upwelling and to support the high level of plankton production. Nutrient fluxes from
mixing, probably more so in winter and spring than summer and fall, offshore, rather than atmospheric, groundwater, volcanogenic, or even
when a seasonal thermocline would have been weakened by atmo- fluvial sources, drove primary production, similar to conditions in the
spheric cooling. Their mean rate of input suggests that the RT of ocean today. Whereas local nutrient fluxes to the photic layer of the
bottom water was a few years to decades. Although much lower than ocean are achieved by vertical mixing and upwelling from shallow sub-
the RT in modern anoxic basins, it suggests seasonal overturns of surface depths, these nutrients are transported from high southern
bottom water were unlikely. latitudes at sub-thermocline depths throughout the oceans. The
90 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

nutrients supplied by this “conveyor” are now and were in Late Jurassic Anderson, R.F., Fleisher, M.Q., Lao, Y., 2006. Glacial–interglacial variability in the
delivery of dust to the central equatorial Pacific Ocean. Earth and Planetary Science
time ultimately derived from the deep Atlantic and or Arctic, wherein Letters 242, 406–414.
they accumulated from deep water particle recycling. This process Anderson, R.F., Fleisher, M.Q., LeHuray, A.P., 1989. Concentration, oxidation state, and
effectively recycles and re-routes nutrients delivered to the global particulate flux of uranium in the Black Sea. Geochimica et Cosmochimica Acta 53,
2215–2224.
ocean from their ultimate sources, such as fluvial inputs from the Arthur, M.A., Dean, W.E., 1999. Organic-matter production and preservation and
continents and from the global atmosphere (Sarmiento and Gruber, evolution of anoxia in the Holocene Black Sea. Paleoceanogaphy 13, 395–411.
2006). Waters with such nutrient inventories currently flow into the Arthur, M.A., Dean, W.E., in press. Oceanographic controls on sedimentary and
geochemical facies on the Peru outer shelf and upper slope. Geological Society of
Cariaco Basin across the shallow sill separating the basin from the deep America Bulletin.
Caribbean to the north. In the case of the Black Sea, nutrients and Arthur, M.A., Sageman, B.B., 1994. Marine black shales: depositional mechanisms and
metals are supplied to the basin in the sub-surface Mediterranean flow environments of ancient deposits. Annual Reviews of Earth and Planetary Science
22, 499–551.
through the Bosporus Strait, which gains nutrients during its passage
Arthur, M.A., Dean, W.E., Hay, B.T., Neff, E.D., King, J., 1994. Varve calibrated records of
through the Aegean Sea and the eastern Mediterranean. Although the carbonate and organic carbon accumulation over the last 2000 years in the Black
nutrient concentrations in these shallow sub-surface waters are lower Sea. Global Biogeochemical Cycles 8, 195–217.
than at abyssal depths of the ocean, they are high enough to dwarf Arthur, M.A., Dean, W.E., Laarkamp, K., 1998. Organic carbon accumulation and
preservation in surface sediments on the Peru margin. Chemical Geology 152,
nutrient supplies from local rivers, owing to the volumes of flow. A 273–286.
similar mechanism must have operated in the Kimmeridge Sea to Bacon, M.P., Brewer, P.G., Spencer, D.W., Murray, J.W., Goddard, J., 1980. Lead-210,
account for the organic enrichment in the KCF, where exceptionally polonium-210, manganese and iron in the Caraico Trench. Deep-Sea Research 27A,
119–135.
high levels of primary production extended virtually uninterruptedly Bailey, G.W., 1991. Organic carbon flux and development of oxygen deficiency on the
over time periods possibly as long as 41,000 years. modern Benguela continental shelf south of 22°S: spatial and temporal variability.
The combined evidence for this advection scheme, together with an In: Tyson, R.V., Pearson, T.H. (Eds.), Modern and Ancient Continental Shelf Anoxia.
The Geological Society (London), Special Publications, vol. 58, pp. 171–183.
elevated sea level in Late Jurassic, suggests that the deposition of the Baird, G.C., Brett, C.E., 1991. Submarine erosion on the anoxic seafloor: stratinomic,
KCF occurred during and as a result of a major global flooding event. palaeoenvironmental, and temporal significance of reworked pyrite bone deposits.
Sea level may have reached several tens of meters above present, In: Tyson, R.V., Pearson, T.H. (Eds.), Modern and Ancient Continental Shelf Anoxia.
Geological Society (London), Special Publications 58, 233–257.
attaining its maximum at the beginning of deposition of ORI-4, Barber, R.T., Smith, R.L., 1981. Coastal upwelling ecosystems. In: Longhurst, A.R. (Ed.),
although the high-stand may have persisted throughout deposition of Analysis of Marine Ecosystems. Academic Press, London, pp. 31–68.
most of the ORIs and the intervening OPIs. Such flooding could have Behrenfeld, M.J., Falkowski, P.G., 1997. Photosynthetic rates derived from satellite-based
chlorophyll concentration. Limnology and Oceanography 42, 1–20.
pushed the major surface currents of the open ocean onto the
Berger, W.H., Johnson, T.C., 1976. Deep-sea carbonates: dissolution and mass wasting on
continents. Today, these currents are restricted to the continental Ontong Java Plateau. Science 192, 785–787.
margins, where Ekman transport of surface waters away from the Berger, W.H., Fischer, K., Lai, C., Wu, G., 1988. Ocean carbon flux: global maps of primary
coasts supports high primary productivity. Their incursion onto the production and export production. In: Agegian, C.R. (Ed.), Biogeochemical Cycling
and Fluxes between the Deep Euphotic Zone and Other Oceanic Realms. NOAA
continents in the Late Jurassic would have transferred into epeiric seas National Undersea Research Program, Report 88-1, pp. 131–176.
the enormous inventory of nutrients they transport between 50 and Berner, R.A., 1970. Sedimentary pyrite formation. American Journal of Science 268, 1–23.
250 m depth and from which upwelling into the photic zone occurs. Berner, R.A., 1981. A new geochemical classification of sedimentary environments.
Journal of Sedimentary Petrology 51, 359–365.
The sea level high-stand in Late Jurassic, more so than the actual sea- Berner, R.A., 1982. Burial of organic carbon and pyrite sulfur in the modern ocean: its
level rise itself, thus was a critical factor in the accumulation of Corg geochemical and environmental significance. American Journal of Science 282,
enriched sediments in the Kimmeridge Sea. Flooding of large areas of 451–473.
Berry, W.B.N., Wilde, P., Quinby-Hunt, M.S., 1989. Paleozoic (Cambrian through Devonian)
low relief in northern and western Europe by as little as a 50 m sea- anoxitropic biotopes. Palaeogeography Palaeoclimatology Palaeoecology 74, 3–13.
level rise would further have resulted in trapping much of the Bertrand, P., Lallier-Verges, E., Boussafir, M., 1994. Enhancement of accumulation and
terrestrial sediment in river estuaries. The persistence of such an anoxic degradation of organic matter controlled by cyclic productivity: a model.
Organic Geochemistry 22, 511–520.
environment at a single site for thousands to several million years Betts, J.N., Holland, H.D., 1991. The oxygen content of bottom waters, the burial
(Herbin et al., 1995) was maintained by the lateral extension of the efficiency of organic carbon, and the regulation of atmospheric oxygen. Palaeogeo-
environment over tens of thousands to several hundred thousand graphy Palaeoclimatology Palaeoecology 97, 5–18.
Bjerrum, C.J., Surlyk, F., Callomon, J.H., Slingerland, R.L., 2001. Numerical paleoceano-
square kilometers, from the Barents Sea to the English Channel. Thus,
graphic study of the Early Jurassic transcontinental Laurasian Seaway (Paper
the oceanographic conditions responsible for the accumulation of the 2000PA000512). Paleoceanography 16, 390–404.
organic-rich Kimmeridge Clay facies, obtained from the modeling Blakey, R., 2001. Global plate tectonics and paleogeography. http://jan.ucc.nau.edu/~rcb7/.
results at a single core site, are likely representative of this vast area. Boesen, C., Postma, D., 1988. Pyrite formation in anoxic environments of the Baltic.
American Journal of Science 288, 575–603.
Bonatti, E., Fisher, D.E., Joenssu, O., Rydell, H.S., 1971. Postdepositional mobility of some
Acknowledgements transition elements, phosphorus, uranium and thorium in deep-sea sediments.
Geochimica et Cosmochimica Acta 35, 189–201.
Nicolas Tribovillard generously provided the chemical data for the Bonilla, J., Senior, W., Bugden, J., Zafiriou, O., Jones, R., 1993. Seasonal distribution of
nutrients and primary productivity on the eastern continental shelf of Venezuela as
Kimmeridge Clay Formation and reviewed an early draft of the influenced by the Orinoco River. Journal of Geophysical Research 98, 2245–2257.
manuscript. Reviews by James Hein, Walter Dean, Richard Tyson, and Boyle, E.A., Sclater, F., Edmond, J.M., 1976. On the marine geochemistry of cadmium.
an anonymous reviewer also greatly improved the manuscript. Nature 263, 42–44.
Boyle, E.A., Sclater, F.R., Edmond, J.M., 1977. The distribution of dissolved copper in the
Support for this research was provided by the Natural Sciences and
Pacific. Earth and Planetary Science Letters 37, 38–54.
Engineering Research Council of Canada (SEC) and the United States Breck, W.G., 1974. Redox levels in the sea. In: Goldberg, E.D. (Ed.), The Sea, vol. 5. Wiley,
Geological Survey (DZP). New York, pp. 153–179.
Brewer, P.G., Spencer, D.W., 1974. Distribution of some trace elements in Black Sea and
their flux between dissolved and particulate phases. In: Degens, E.T., Ross, D.A.
References (Eds.), The Black Sea—Geology, Chemistry, and Biology. Memoir, vol. 20. American
Association of Petroleum Geologists, Tulsa, pp. 137–143.
Algeo, T.J., 2004. Can marine anoxic events draw down the trace element inventory of Brink, K.H., Jones, B.H., Van Leer, J.C., Mooers, C.N.K., Stuart, D.W., Stevenson, M.R., Dugdale,
seawater? Geology 32, 1057–1060. R.C., Heburn, G.W., 1981. Physical and biological structure and variability in an
Aller, R.C., Blair, N.E., 1996. Sulfur diagenesis and burial on the Amazon shelf: major upwelling center off Peru near 15 degrees during March, 1977. In: Richards, F.A. (Ed.),
control by physical sedimentation processes. Geo-Marine Letters 16, 3–10. Coastal Upwelling, vol. I. American Geophysical Union, Washington, DC, pp. 473–495.
Aller, R.C., Blair, N.E., Xia, Q., Rude, P.D., 1996. Remineralization rates, recycling, and storage Brock, J.C., McClain, C.R., 1992. Interannual variability in phytoplankton blooms
of carbon in Amazon shelf sediments. Continental Shelf Research 16, 753–786. observed in the northwestern Arabian Sea during the southwest monsoon. Journal
Anderson, R.F., Fleisher, M.Q., 1991. Uranium precipitation in Black Sea sediments. In: of Geophysical Research 97, 733–750.
Izdar, E., Murray, J.W. (Eds.), Black Sea Oceanography. NATO ASI Series. Kluwer Brockamp, B., 1944. Zur Pälogeographie und Bitumenführung des Posidonienschiefers
Academic Publishers, Dordrecht, Netherlands, pp. 443–458. im deutschen Lias. Archiv für Lagerstättenforschung 77, 7–59.
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 91

Brüchert, V., Jørgensen, B.B., Neumann, K., Riechmann, D., Schlösser, M., Schulz, H., Collier, R.W., Edmond, J.M., 1983. Plankton composition and trace element fluxes from
2003. Regulation of bacterial sulfate reduction and hydrogen sulfide fluxes in the the surface ocean. In: Wong, C.S., Boyle, E., Bruland, K.W., Burton, J.D., Goldberg, E.D.
central Namibian coastal upwelling zone. Geochimica et Cosmochimica Acta 67, (Eds.), Trace Metals in Seawater, NATO Conference Series IV-9. Plenum Press,
4505–4518. New York, pp. 789–809.
Bruland, K.W., 1983. Trace elements in sea-water, In: Riley, J.P., Skirrow, G. (Eds.), 2nd Condie, K.C., 1992. Chemical composition and evolution of the upper continental crust:
Edition. Chemical Oceanography, vol. 8. Academic Press, London, pp. 157–220. contrasting results from surface sediments and shales. Chemical Geology 104, 1–37.
Brumsack, H.J., 1986. The inorganic geochemistry of Cretaceous black shales (DSDP Leg 41) Coote, A.R., Yeats, P.A., 1979. Distribution of nutrients in the Gulf of St. Lawrence. Journal
in comparison to modern upwelling sediments from the Gulf of California. In: of the Fisheries Research Board of Canada 36, 122–131.
Summerhayes, C.P., Shackleton, N.J. (Eds.), North Atlantic Palaeoceanography. Cope, J.C.W., 1974. New information of the Kimmerdige Clay of Yorkshire. Proceedings
Geological Society Special Publications, vol. 21. Geological Society of London, London, Geologists Association 85, 211–221.
United Kingdom, pp. 447–462. Copenhagen, W.J., 1953. The periodic mortality of fish in the Walvis region: a phenomenon
Buggisch, W., 1991. The global Frasnian-Famennian “Kellwasser Event”. Geologische within the Benguela Current. Investigational Report, Division of Fisheries, South Africa
Rundschau 80, 49–72. 14, 1–35.
Burnett, W.C., Bokuniewicz, H., Huettel, M., Moore, W.S., Taniguchi, M., 2003. Corliss, J.B., Dymond, J., Gordon, L.I., Edmond, J.M., Von Herzen, R.P., Ballard, R.D., Green,
Groundwater and pore water inputs to the coastal zone. Biogeochemistry 66, 3–33. K., Williams, D., Bainbridge, A., Crane, K., van Andel, T.H., 1979. Submarine thermal
Burnett, W.C., Froelich, P.N., 1988. The origin of marine phosphorite—the results of the springs on the Galapagos Rift. Science 203, 1073–1083.
R/V Robert D. Conrad Cruise 23-06 to the Peru shelf. Marine Geology 80, 181–343. Corredor, J., Morell, J., Lopez, J., Armstrong, R., Dieppa, A., Cabanillas, C., Cabrera, A.,
Byers, C.W., 1977. Biofacies patterns in euxinic basins: a general model. In: Cook, H.E., Hensley, V., 2003. Remote continental forcing of phytoplankton biogeochemistry:
Enos, P. (Eds.), Deep-water Carbonate Environments. Special Publication, vol. 25. observations across the “Caribbean–Atlantic front”. Geophysical Research Letters
Society of Economic Paleontologists and Mineralogists, Tulsa, OK, pp. 5–17. 30. doi:10.1029/2003GL018193. OCE 4.
Cadée, G.C., 1978. Primary production and chlorophyll in the Ziare River, estuary and Cotté-Krief, M.-H., Thomas, A.J., Martin, J.-M., 2002. Trace metal (Cd, Cu, Ni, and Pb)
plume. Netherlands Journal of Sea Research 12, 368–381. cycling in the upper water column near the shelf edge of the European continental
Calvert, S.E., 1964. Factors affecting distribution of laminated diatomaceous sediments margin (Celtic Sea). Marine Chemistry 79, 1–26.
in Gulf of California. In: Van Andel, T.J., Shor, G.G. (Eds.), Marine Geology of the Gulf Cox, B.M. and Gallois, R.W., 1981. The stratigraphy of the Kimmeridge Clay of the Dorset
of California. Memoir, vol. 3. American Association of Petroleum Geologists, Tulsa, type area and its correlation with some other Kimmeridgian sequences. Institute of
pp. 311–330. Geological Sciences, Report 80/4, London.
Calvert, S.E., 1990. Geochemistry and origin of the Holocene sapropel in the Black Sea. Crusius, J., Anderson, R.F., 1992. Inconsistencies in accumulation rates of Black Sea
In: Ittekkot, V., Kempe, S., Michaelis, M., Spitzy, A. (Eds.), Facets of Modern sediments inferred from records of laminae and 210Pb. Paleoceanography 7, 215–227.
Biogeochemistry. Springer-Verlag, Berlin, pp. 326–352. Crusius, J., Calvert, S.E., Pedersen, T.F., Sage, D., 1996. Rhenium and molybdenum
Calvert, S.E., Karlin, R.E.,1991. Relationships between sulphur, organic carbon, and iron in the enrichments in sediments as indicators of oxic, suboxic and anoxic conditions of
modern sediments of the Black Sea. Geochimica et Cosmochimica Acta 55, 2483–2490. deposition. Earth and Planetary Science Letters 145, 65–78.
Calvert, S.E., Pedersen, T.F., 2007. Elemental proxies for palaeoclimatic and palaeoceano- Dalsgaard, T., Canfield, D.E., Petersen, J., Thamdrup, B., Acuna-Gonzalez, J., 2003. N2
graphic variability in marine sediments: interpretation and application. In: Hillaire- production by the anammox reaction in the anoxic water column of Golfo Dulce,
Marcel, C., de Vernal, A. (Eds.), Paleoceanography of the Late Cenozoic. Part 1. Methods Costa Rica. Nature 422, 606–607.
in Late Cenozoic Paleoceanography. Elsevier, New York, pp. 567–644. Dean, W.E., Gardner, J.V., 1998. Pleistocene to Holocene contrasts in organic matter
Calvert, S.E., Price, N.B., 1971. Upwelling and nutrient regeneration in the Benguela production and preservation on the California continental margin. Geological
Current, October, 1968. Deep-Sea Research 18, 505–523. Society of America Bulletin 110, 888–899.
Calvert, S.E., Price, N.B., 1983. Geochemistry of Namibian Shelf sediments. In: Suess, E., DeBaar, H.J.W., German, C.R., Elderfield, H., van Gaans, P., 1988. Rare earth element
Thiede, J. (Eds.), Coastal Upwelling, NATO Conference Series IV-10a. Plenum distributions in anoxic waters of the Cariaco Trench. Geochimica et Cosmochimica
Publishing Corporation, New York, pp. 337–375. Acta 52, 1203–1219.
Calvert, S.E., Karlin, R.E., Toolin, L.J., Donahue, D.J., Southon, J.R., Vogel, J.S., 1991. Low Degens, E.T., Ross, D.A., 1972. Chronology of the Black Sea over the last 25,000 years.
organic carbon accumulation rates in Black Sea sediments. Nature 350, 692–694. Chemical Geology 10, 1–16.
Calvert, S.E., Pedersen, T.F., Karlin, R.E., 2001. Geochemical and isotopic evidence for Dehaene, S., Izard, V., Spelke, E., Pica, P., 2008. Log or linear? Distinct intuitions of the
post-glacial palaeoceanographic changes in Saanich Inlet, British Columbia. Marine number scale in western and Amazonian indigene cultures. Science 320, 1217–1220.
Geology 174, 287–305. Demaison, G.J., Moore, G.T., 1980. Anoxic environments and oil source bed genesis.
Canfield, D., 1989. Sulfate reduction and oxic respiration in marine sediments: Organic Geochemistry 2, 9–31.
implication for organic carbon preservation in euxinic environments. Deep-Sea DeMaster, D.J., Pope, R.H., 1996. Nutrient dynamics in Amazon shelf waters: results from
Research 36, 121–138. AMASSEDS. Continental Shelf Research 16, 263–289.
Canfield, D.E., 1994. Factors influencing organic carbon preservation in marine DeMaster, D.J., Smith, W.O., Nelson, D.M., Aller, J.Y., 1996. Biogeochemical processes in
sediments. Chemical Geology 114, 315–329. Amazon shelf waters: chemical distributions and uptake rates of silicon, carbon and
Caplan, M.L., Bustin, R., 2001. Palaeoenvironmental and palaeoceanographic controls on nitrogen. Continental Shelf Research 16, 617–643.
black, laminated mudrock deposition: example from Devonian–Carboniferous Deuser, W.G., 1973. Cariaco Trench: oxidation of organic matter and residence time of
strata, Alberta, Canada. Sedimentary Geology 145 (ER1-2), 45–72. anoxic water. Nature 242, 601–603.
Capone, D.G., Bautista, M.F., 1985. A groundwater source of nitrate in nearshore marine Deuser, W.G., 1974. Evolution of anoxic conditions in Black Sea during Holocene. In:
sediments. Nature 313, 214–216. Degens, E.T., Ross, D.A. (Eds.), The Black Sea—Geology, Chemistry, and Biology.
Caspers, H., 1957. Black Sea and Sea of Azov. In: Hedgpeth, J.W. (Ed.), Treatise on Marine Memoir, vol. 20. American Association of Petroleum Geologists, Tulsa, pp. 133–136.
Ecology. Geological Society of America, pp. 803–890. Dickson, M.L., Orchardo, J., Barber, R.T., Marra, J., McCarthy, J.J., Sambrotto, R.N., 2001.
Chavez, F.P., Barber, R.T., 1987. An estimate of new production in the equatorial Pacific. Production and respiration rates in the Arabian Sea during the 1995 northeast and
Deep-Sea Research 34, 1229–1243. southwest monsoons. Deep Sea Research 48, 1199–1230.
Chavez, F.P., Smith, S.L., 1995. Biological and chemical consequences of open ocean Doty, M.S., Oguri, M., 1956. The island mass effect. Journal du Conseil Permanent
upwelling. In: Summerhayes, C.P., Emeis, K.-C., Angel, M.V., Smith, R.L., Zeitzschel, B. International pour l'Exploration de la Mer 22, 33–37.
(Eds.), Upwelling in the Ocean: Modern Processes and Ancient Records. J. Wiley, Douglas, R., Gonzalez Yajimovich, O., Ledesma Vazquez, J., Staines Urias, F., 2007. Climate
New York, pp. 149–169. forcing, primary production and the distribution of Holocene biogenic sediments in
Chavez, F.P., Toggweiler, J.R., 1995. Physical estimates of global new production: the the Gulf of California. Quaternary Science Reviews 26, 115–129.
upwelling contribution. In: Summerhayes, C.P., Emeis, K.-C., Angel, M.V., Smith, R.L., Dugdale, R.C., Goering, J.J., Barber, R.T., Smith, R.L., Packard, T.T., 1977. Denitrification and
Zeitzschel, B. (Eds.), Upwelling in the Ocean. Modern Processes and Ancient hydrogen sulfide in the Peru upwelling region during 1976. Deep Sea Research 24,
Records. J. Wiley, New York, pp. 313–320. 601–608.
Chester, R., Hughes, M.J., 1967. A chemical technique for the separation of ferromanga- Dyrssen, D., 1985. Metal complex formation in sulphidic seawater. Marine Chemistry 15,
nese minerals, carbonate minerals and adsorbed trace elements from pelagic 285–293.
sediments. Chemical Geology 2, 249–262. Edmond, J.M., Boyle, E.A., Grant, B., Stallard, R.F., 1981. The chemical mass balance in the
Codispoti, L.A., 1980. Temporal nutrient variability in three different upwelling regimes. Amazon plume I: the nutrients. Deep-Sea Research 28A, 1339–1374.
In: Richards, F.A. (Ed.), Coastal Upwelling, vol. I. American Geophysical Union, Edmond, J.M., Spivack, A., Grant, B.C., Ming-Hui, H., Zexiam, C., Sung, C., Xiushau, Z., 1985.
Washington, D.C., pp. 209–220. Chemical dynamics of the Changjiang Estuary. In: Milliman, J.D., Jin, Q. (Eds.), Sediment
Codispoti, L.A., 1989. Phosphorus vs nitrogen limitation of new and export production. Dynamics of the Changjiang Estuary and the Adjacent East China Sea. Continental Shelf
In: Berger, W.H., Smetacek, V.S., Wefer, G. (Eds.), Productivity of the Ocean: Present Research, vol. 4 (1–2). Pergamon, Oxford, United Kingdom, pp. 17–36.
and Past. J. Wiley, New York, pp. 377–394. Elderfield, H., Greaves, M.J., 1982. The rare earth elements in seawater. Nature 296,
Codispoti, L.A., Brandes, J.A., Christensen, J.P., Devol, A.H., Naqvi, S.W.A., Paerl, H.W., 214–219.
Yoshinari, T., 2001. The oceanic fixed nitrogen and nitrous oxide budgets: moving Ellwood, B.B., Ledbetter, M.T., 1979. Paleocurrent indicators in deep-sea sediments.
targets as we enter the anthropocene? Scientia Marina 65 (supplement 2), 85–105. Science 203, 1335–1337.
Codispoti, L.A., Friederich, G.E., Murray, J.W., Sakamoto, C.M., 1991. Chemical variability Emerson, S.R., Huested, S.S., 1991. Ocean anoxia and the concentrations of molybdenum
in the Black Sea: implications of continuous vertical profiles that penetrated the and vanadium in seawater. Marine Chemistry 34, 177–196.
oxic/anoxic interface. Deep-Sea Research 38, S691–S710. Eppley, R., Peterson, B.J., 1979. Particulate organic matter flux and planktonic new
Collier, R.W., 1984. Particulate and dissolved vanadium in the North Pacific Ocean. production in the deep ocean. Nature 282, 677–680.
Nature 309, 441–444. Ettensohn, F.R., 1995. Global and regional controls on the origin and burial of organic
Collier, R.W., 1985. Molybdenum in the northeast Pacific Ocean. Limnology and matter in Devonian–Mississippian black shales of North America. Houston Geological
Oceanography 30, 1351–1354. Society Bulletin 37, 12–17.
92 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

Fairbanks, R.G., 1989. A 17,000-year glacio-eustatic sea level record: influence of glacial Gromet, P.L., Dymek, P.F., Haskin, L.A., Korotev, R.L., 1984. The “North American shale
melting rates on the Younger Dryas event and deep-ocean circulation. Nature 342, composite”: its composition, major and minor element characteristics. Geochimica
637–642. et Cosmochimica Acta 48, 2469–2482.
Fanning, K.A., Pilson, M.E.Q., 1972. A model for the anoxic zone of the Cariaco Trench. Gustafson, L.B., Williams, N., 1981. Sediment-hosted stratiform deposits of copper, lead,
Deep-Sea Research 19, 847–863. and zinc. Economic Geology 75th Anniversary Volume, pp. 139–178.
Fonseca, C., 2000. Upwelling, sedimentation and anoxia control on deposition of Halim, Y., 1991. The impact of human alterations of the hydrological cycle on ocean
phosphates in the Late Cretaceous California margin. In: Glenn, C.R., Prévôt-Lucas, L., margins. In: Mantoura, R.F.C., Martin, J.-M., Wollast, R. (Eds.), Ocean-margin
Lucas, J. (Eds.), Marine Authigenesis: From Global to Microbial.. Special Publication, Processes in Global Change. John Wiley and Sons, New York, pp. 301–327.
vol. 66. Society for Economic Paleontology and Mineralogy, Tulsa, OK, pp. 455–480. Hallam, A., 1975. Jurassic Environments. Cambridge University Press, Cambridge. 269 pp.
Ford, T.E., Naiman, R.J., 1989. Groundwater–surface water relationships in boreal forest Hallam, A., Bradshaw, M.J., 1979. Bituminous shales and oolitic ironstones as indicators of
watersheds: dissolved organic carbon and inorganic nutrient dynamics. Canadian transgressions and regressions. Journal of the Geological Society, London 136,157–164.
Journal of Fisheries and Aquatic Sciences 46, 41–49. Hampel, C.A. (Ed.), 1968. The encyclopedia of the chemical elements. Reinhold Book
François, R., 1988. A study on the regulation of the concentrations of some trace metals Corporation, New York. 849 pp.
(Rb, Sr, Zn, Pb, Cu, V, Cr, Ni, Mn and Mo) in Saanich Inlet sediments, British Haq, B.U., Hardenbol, J., Vail, P.R., 1987. Chronology of fluctuating sea levels since the
Columbia, Canada. Marine Geology 83, 285–308. Triassic. Science 235, 1156–1166.
François, R., Bacon, M.P., Suman, D.O., 1990. Thorium 230 profiling in deep sea Haraldsson, C., Westerlund, S., 1988. Trace metals in the water columns of the Black Sea
sediments: high-resolution records of flux and dissolution of carbonate in the and Framvaren Fjord. Marine Chemistry 23, 417–424.
equatorial Atlantic during the last 24,000 years. Paleoceanography 5, 761–787. Haraldsson, C., Westerlund, S., 1991. Total and suspended cadmium, cobalt, copper, iron, lead,
François, R., Frank, M., van der Loeff, M.M.R., Bacon, M.P., 2004. 230Th normalization: an manganese, nickel and zinc in the water column of the Black Sea. In: Murray, J.W., Izdar, E.
essential tool for interpreting sedimentary fluxes during the late Quaternary. (Eds.), Black Sea Oceanography. Kluwer, Dordrecht, pp. 161–172. NATO ASI Series.
Paleoceanography 19, PA1018. doi:10.1029/2003PA000939. Harris, N.B., 2005. The deposition of organic-carbon-rich sediments: models, mechanisms,
Frank, M., Eisenhauer, A., Bonn, W.J., Walter, P., Grobe, H., Kubik, P.W., Dittrich-Hannen, and consequences—introduction. In: Harris, N.B. (Ed.), The Deposition of Organic-
B., Mangini, A., 1995. Sediment redistribution versus paleoproductivity change: carbon-rich Sediments: Models, Mechanisms, and Consequences. Special Publication,
Weddell Sea margin sediment stratigraphy and biogenic particle flux of the last vol. 82. Society for Sedimentary Geology, Tulsa, OK, pp. 1–5.
250,000 years deduced from 230Thex, 10Be and biogenic barium profiles. Earth and Hatch, J.R., Leventhal, J.S., 1992. Relationship between inferred redox potential of the
Planetary Science Letters 136, 559–573. depositional environment and geochemistry of the Upper Pennsylvanian (Mis-
Galeotti, S., Sprovieri, M., Coccioni, R., Bellanca, A., Neri, R., 2003. Orbitally modulated black sourian) Stark Shale Member of the Dennis Limestone, Wabaunsee County, Kansas,
shale deposition in the upper Albian Amadeus Segment (central Italy): a multi-proxy U.S.A. Chemical Geology 99, 65–82.
reconstruction. Palaeogeography Palaeoclimatology Palaeoecology 190, 441–458. Haug, G.H., Hughen, K.A., Sigman, D.M., Peterson, L.C., Rohl, U., 2001. Southward migration
Ganeshram, R.S., Pedersen, T.F., 1998. Glacial–interglacial variability in upwelling and of the Intertropical Convergence Zone through the Holocene. Science 293, 1304–1307.
bioproductivity off NW Mexico: implication for Quaternary paleoclimate. Paleo- Hebting, Y., Schaeffer, P., Behrens, A., Adam, P., Schmitt, G., Schneckenburger, P.,
ceanography 13, 634–645. Bernasconi, S.M., Albrecht, P., 2006. Biomarker evidence for a major preservation
Garrels, R.M., Christ, C.L., 1965. Solutions, Minerals, and Equilibria. Freeman, Cooper & pathway of sedimentary organic carbon. Science 312, 1627–1630.
Company, San Francisco. 450 pp. Hedges, J.I., Keil, R.G., 1995. Sedimentary organic matter preservation: an assessment
German, C.R., Holliday, B.P., Elderfield, H., 1993. Reply to the comment by E.R. and speculative synthesis. Marine Chemistry 49, 81–115.
Sholkovitz, on “Redox cycling of rare earth elements in the suboxic zone of the Black Hedges, J.I., Baldock, J.A., Gellnas, Y., Lee, C., Peterson, M., Wakeman, S.G., 2001. Evidence
Sea”. Geochimica et Cosmochimica Acta 56, 4309–4313. of non-selective preservation of organic matter in sinking marine particles. Nature
Geyer, W.R., Beardsley, R.C., Lentz, S.J., Candela, J., Limebumer, R., Johns, W.E., Castro, B.M., 409, 801–804.
Soares, I.D., 1996. Physical oceanography of the Amazon shelf. Continental Shelf Helz, G.R., Miller, C.V., Charnock, J.M., Mosselmans, J.F.W., Pattrick, R.A.D., Garner, C.D.,
Research 16, 575–616. Vaughan, D.J., 1996. Mechanism of molybdenum removal from the sea and its
Gibbs, R.J., 1967. The geochemistry of the Amazon River system: Part I. The factors that concentration in black shales: EXAFS evidence. Geochimica et Cosmochimica Acta
control the salinity and the composition and concentration of the suspended solids. 60, 3631–3642.
Geological Society of America Bulletin 78, 1203–1232. Henrichs, S.M., Reebugh, W.S., 1987. Anaerobic mineralization of marine sediment
Gibbs, R.J., 1980. Wind-controlled coastal upwelling in the western equatorial Atlantic. organic matter: rates and role of anaerobic processes in the oceanic carbon
Deep-Sea Research 27A, 857–866. economy. Geomicrobiology Journal 5, 191–237.
Giblin, A.E., Gaines, A.G., 1990. Nitrogen inputs to a marine embayment: the importance Herbin, J.P., Geyssant, J.R., 1993. “Ceintures organiques” au Kimmeridgien/Tithonien en
of groundwater. Biogeochemistry 10, 309–328. Angleterre (Yorkshire, Dorset) et en France (Boulonnais). Comptes Redus de
Glazer, B.T., Luther III, G.W., Konovalov, S.K., Friederich, G.E., Nuzzio, D.B., Trouwborst, R.E., l'Académie des Sciences 317, 1309–1316.
Tebo, B.M., Clement, B., Murray, K., Romanov, A.S., 2006. Documenting the suboxic Herbin, J.P., Fernandez-Martinez, L.J., Geyssant, J.R., Albani, A.E., Deconinck, J.F., Proust, J.N.,
zone of the Black Sea via high-resolution real-time redox profiling. Deep Sea Research Colbeaux, J.P., Vidier, J.P., 1995. Sequence stratigraphy of source rocks applied to the
53, 1740–1755. study of the Kimmeridge/Tithonian in the Northwest European Shelf (Dorset/UK,
Glenn, C.R., Arthur, M.A., 1984. Sedimentary and geochemical indicators of productivity Yorkshire/UK, and Boulonnais/France). Marine and Petroleum Geology 12, 177–194.
and oxygen contents in modern and ancient basins: the Holocene Black Sea as the Heywood, K.J., Barton, E.D., Simpson, J.H., 1990. The effects of flow disturbance by an
“type” anoxic basin. Chemical Geology 48, 325–354. oceanic island. Journal of Marine Research 48, 55–73.
Goldberg, E.D., 1963. Mineralogy and chemistry of marine sedimentation. In: Shepard, F.P. Hidalgo Gonzalez, R.M., Alvarez Borrego, S., 2004. Total and new production in the Gulf
(Ed.), Submarine Geology. Harper and Row, New York, pp. 436–466. of California estimated from ocean color data from the satellite sensor SeaWIFS.
Goldhaber, M.B., Kaplan, I.R., 1974. The sulfur cycle. In: Goldberg, E.D. (Ed.), The Sea, vol. 5. Deep-Sea Research 51, 739–752.
John Wiley, New York, pp. 569–655. Hirst, D.M., 1974. Geochemistry of sediments from eleven Black Sea cores. In: Degens, E.T.,
Goldhaber, M.B., Aller, R.C., Cochran, J.K., Rosenfeld, J.K., Martens, C.S., Berner, R.A., 1977. Ross, D.A. (Eds.), The Black Sea—Geology, Chemistry and Biology. American
Sulfate reduction, diffusion, and bioturbation in Long Island Sound sediments: Association of Petroleum Geologists Memoir, vol. 20, pp. 430–455. Tulsa, OK.
report of the FOAM group. American Journal of Science 177, 193–237. Ho, T.Y., Quigg, A., Finkel, Z.V., Milligan, A.J., Wyman, K., Falkowski, P.G., Morel, F.M.,
Goñi, M.G., Heather, L.A., Thunell, R.C., Tappa, E., Black, D., Astor, Y., Varela, R., Müller- 2003. The elemental composition of some marine phytoplankton. Journal of
Karger, F., 2003. Biogenic fluxes in the Cariaco Basin: a combined study of sinking Phycology 39, 1145–1159.
particulates and underlying sediments. Deep-Sea Research 50, 781–807. Honjo, S., Hay, B.J., Manganini, S.J., Asper, V.L., Degens, E.T., Kempe, S., Ittekkot, V., Izdar, E.,
Gordon, A.L., 1971. Oceanography of Antarctic waters. In: Reid, J.L. (Ed.), Antarctic Konuk, T., Benli, H.A., 1987. Seasonal cyclicity of lithogenic particle fluxes at a southern
Oceanology I. American Geophysical Union, Washington, pp. 169–203. Black Sea sediment trap station. In: Degens, E.T., Izdar, E., Honjo, S. (Eds.), Particle Flux
Gong, G.-C., Wen, Y.-H., Wang, B.-W., Liu, G.-J., 2003. Seasonal variation of chlorophyll a in the Ocean. SCOPE/UNEP Sonderband, Number 62. University of Hamburg,
concentration, primary production and environmental conditions in the subtropical Hamburg, pp. 19–39.
East China Sea. Deep-Sea Research 50, 1219–1236. House, M.R., 1985. A new approach to the absolute time scale from measurements of
Goyet, C., Bradshaw, A.L., Brewer, P.G., 1991. The carbonate system in the Black Sea. orbital cycles and sedimentary microrhythms. Nature 315, 721–725.
Deep-Sea Research 38, S1049–S1068. Hseuh, Y., O'Brien, J.J., 1971. Steady coastal upwelling induced by an along-shore current.
Gradstein, F.M., Agterberg, F.P., Ogg, J.G., Hardenbol, J., van Veen, P., Thierry, J., Zehul, H., Journal of Physical Oceanography 1, 180–186.
1995. A Triassic, Jurassic, and Cretaceous time scale. In: Kent, W.A., Aubry, M.-P., Huc, A.Y., Lallier-Verges, E., Bertrand, P., Carpentier, B., Hollander, D.J., 1992. Organic
Hardenbol, J. (Eds.), Geochronology, Time Scales and Global Statigraphic Correla- matter response to change of depositional environment in Kimmeridgian Shales,
tions. Society of Economic Paleontologists and Mineralogists, Special Publication, Dorset, U.K. In: Whelan, J.K., Farrington, J.W. (Eds.), Organic Matter: Productivity,
vol. 54, pp. 99–126. Tulsa, OK. Accumulation, and Preservation in Recent and Ancient Sediments. Columbia
Gradstein, F.M., Ogg, J.G., Smith, A.G., 2004. A Geologic Time Scale. Cambridge Press, University Press, New York, pp. 469–486.
Cambridge, p. 589 pp. Hughen, K.A., Overpeck, J.T., Lehman, S.R., Kashgarian, M., Southon, J., Peterson, L.C.,
Graversen, O., 2002. A structural transect between the central North Sea Dome and the Alley, R., Sigman, D.M., 1998. Deglacial changes in ocean circulation from an
South Swedish Dome: Middle Jurassic–Quaternary uplift/subsidence reversal and extended radiocarbon calibration. Nature 391, 65–68.
exhumation across the eastern North Sea Basin. In: Dore, A.G., Cartwright, J., Stoker, Hughen, K., Overpeck, J., Peterson, L., Anderson, R., 1996. The nature of varved
M.S., Turner, M.S., White, N. (Eds.), Examination of the North Atlantic Margin: sedimentation in the Cariaco Basin, Venezuela, and its palaeoclimatic significance.
Timing, Mechanisms and Implications for Petroleum Exploration. The Geological In: Kemp, A.E.S. (Ed.), Palaeoclimatology and Palaeoceanography from Laminated
Society (London), Special Publications, vol. 196, pp. 67–83. Sediments. Geological Society of (London), pp. 171–183.
Grégoire, M.A., Stanev, E.V., 2001. Ventilation of Black Sea anoxic waters. Journal of Huntsman, S.A., Barber, R.T., 1977. Primary productivity off northwest Africa: the
Marine Systems 31, 1–2. relationship to wind and nutrient conditions. Deep-Sea Research 24, 25–33.
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 93

Hutchings, L., Pitcher, G.C., Probyn, T.A., Bailey, G.W., 1995. The chemical and biological Lancelot, C., Rousseau, V., Billen, G., Van Eeckhout, D., 1997. Coastal eutrophication of the
consequences of coastal upwelling. In: Summerhayes, C.P., Emeis, K.-C., Angel, M.V., southern bight of the North Sea: assessment and modelling. Bulletin de la Société
Smith, R.L., Zeitzschel, B. (Eds.), Upwelling in the Ocean. Modern Processes and royale des sciences de Liège 66, 27–38.
Ancient Records. Environmental Science Research Report, vol. 18. J. Wiley, New York, Landing, W.M., Lewis, B.L., 1991. Thermodynamic modeling of trace metal speciation in
pp. 65–81. the Black Sea. In: Murray, J.W., Izdar, E. (Eds.), Black Sea Oceanography. Kluwer,
Hutchins, D.A., Burland, K.W., 1998. Iron-limited diatom growth and Si:N uptake ratios Dordrecht, pp. 125–160. NATO ASI Series.
in a coastal upwelling regime. Nature 393, 561–564. Lane-Serff, G.F., Rohling, E.J., Bryden, H.L., Charnock, H., 1997. Postglacial connection of
Hutchins, D.A., Hare, C.E., weaver, R.S., Zhang, Y., Firme, G.F., DiTullio, G.R., Alm, M.B., the Black Sea to the Mediterranean and its relation to the timing of sapropel
Riseman, S.F., Maucher, J.M., Geesey, M.E., Trick, C.G., Smith, G.J., Rue, L.E., Conn, J., formation. Paleoceanography 12, 169–174.
Bruland, K.W., 2002. Phytoplankton iron limitation in the Humboldt Current and Leckie, D.A., Singh, C., Goodarezi, F., Wall, J.H., 1990. Organic-rich, radioactive marine
Peru Upwelling. Limnology and Oceanography 47, 997–1011. shale: a case study of a shallow-water condensed section, Cretaceous Shaftesbury
Iselin, C.O.D., 1939. Some physical factors which may influence the productivity of New Formation, Alberta, Canada. Journal of Sedimentary Petrology 60, 101–117.
England's coastal waters. Journal of Marine Research 2, 74–85. Lee, B.-S., Bullister, J.I., Murray, J.W., Sonnerup, R.E., 2002. Anthropogenic chlorofluor-
Ivanenkov, V.N., Rozanov, A.G., 1961. Hydrogen sulphide contamination of the ocarbons in the Black Sea and the Sea of Marmara. Deep-Sea Research 49, 895–913.
intermediate water layers of the Arabian Sea and the Bay of Bengal. Oceanology 1, Lees, J.A., Bown, P.R., Young, J.R., Riding, J.B., 2004. Evidence for annual records of
443–449. phytoplankton productivity in the Kimmeridge Clay Formation coccolith stone
Ivanov, L.I., Samodurov, A.S., 2001. The role of lateral fluxes in ventilation of the Black bands (Upper Jurassic, Dorset, UK). Marine Micropalaeontology 52, 29–49.
Sea. Journal of Marine Systems 31, 159–174. Leinen, M., 1977. A normative calculation technique for determining opal in deep-sea
Jacobs, L., Emerson, S., 1982. Trace metal solubility in an anoxic fjord. Earth and sediments. Geochimica et Cosmochimica Acta 41, 671–676.
Planetary Science Letters 60, 237–252. Leventhal, J.S., 1983. An interpretation of carbon and sulphur relationships in Black Sea
Jacobs, L., Emerson, S., Huested, S.S., 1987. Trace metal geochemistry in the Cariaco sediments as indicators of environments of deposition. Geochimica et Cosmochimica
Trench. Deep-Sea Research 34, 965–981. Acta 47, 133–137.
Jacobs, L., Emerson, S., Skei, J., 1985. Partitioning and transport of metals across the O2/ Lewan, M.D., Maynard, J.B., 1982. Factors controlling enrichment of vanadium and nickel
H2S interface in a permanently anoxic basin: Framvaren Fjord, Norway. Geochimica in the bitumen of organic sedimentary rocks. Geochimica et Cosmochimica Acta 46,
et Cosmochimica Acta 49, 1433–1444. 2547–2560.
Jahnke, R.A., Nelson, J.R., Marinelli, R.L., Eckman, J.E., 2000. Benthic flux of biogenic Lewis, B.L., Landing, W.M., 1992. The investigation of dissolved and suspended-particulate
elements on the Southeastern US continental shelf: influence of pore water trace metal fractionation in the Black Sea. Marine Chemistry 40, 105–141.
advective transport and benthic microalgae. Contential Shelf Research 20, 109–127. Lin, H.-L., Peterson, L.C., Overpeck, J.T., Trumbore, S.E., Murray, D.W., 1997. Late
Jahnke, R.A., Nelson, J.R., Richards, M.E., Robertson, C.Y., Rao, A.M.F., Jahnke, D.B., 2008. Quaternary climate change from d18O records of multiple species of planktonic
Benthic primary productivity on the Georgia midcontinental shelf: benthic flux foraminifera: high-resolution records from the anoxic Cariaco Basin, Venezuela.
measurements and high-resolution, continuous in situ PAR records. Journal of Paleoceanography 12, 415–427.
Geophysical Research 113, C08022. doi:10.1029/2008JC004745. Loubere, P., Mekik, F., Francois, R., Pichat, S., 2004. Export fluxes of calcite in the eastern
Jenkyns, H.C., Jones, C.E., Grocke, D.R., Hesselbo, S.P., Parkinson, D.N., 2002. equatorial Pacific from the Last Glacial Maximum to present. Paleoceanography 19.
Chemostratigraphy of the Jurassic System: applications, limitations and implica- doi:10.1029/2003PA000986.
tions for palaeoceanography. Journal Geological Society (London) 159, 351–378. Lyle, M., 1976. Estimation of hydrothermal manganese input to the oceans. Geology 4,
Johannes, R.E., 1980. The ecological significance of the submarine discharge of ground 733–736.
water. Marine Ecology Progress Series 3, 365–373. Lyle, M., 1983. The brown-green color transition in marine sediments: a marker of the
Johnson, D.A., 1972. Ocean-floor erosion in the equatorial Pacific. Geological Society of Fe(III)–Fe(II) redox boundary. Limnology and Oceanography 28, 1026–1033.
America Bulletin 83, 3121–3144. Lyons, T.W., Berner, R.A., 1992. Carbon–sulfur–iron systematics of Upper Holocene Black
Jones, G.A., Gagnon, A.R., 1994. Radiocarbon chronology of Black Sea sediments. Deep- Sea sediments. Chemical Geology 99, 1–27.
Sea Research 41, 531–557. Lyons, T.W., Kashgarian, M., 2005. Paradigm lost, paradigm found: the Black Sea
Jones, B.F., Manning, D.A.C., 1994. Comparison of geochemical indices used for the connection as viewed from the anoxic basin margin. Oceanography 18, 87–99.
interpretation of palaeoredox conditions in ancient mudstones. Chemical Geology Lyons, T.W., Werne, J.P., Hollander, D.J., Murray, R.W., 2003. Contrasting sulfur
111, 111–129. geochemistry and Fe/Al and Mo/Al ratios across the last oxic-to-anoxic transition
Jørgensen, B.B., 1982. Mineralization of organic matter in the sea bed—the role of in the Cariaco Basin, Venezuela. Chemical Geology 195, 131–157.
sulphate reduction. Nature 296, 643–645. Macquaker, J.H.S., Gawthorpe, R.L., 1993. Mudstone lithofacies in the Kimmeridge Clay
Karl, D.M., Knauer, G.A., 1991. Microbial production and particle flux in the upper 350 m Formation, Wessex Basin, southern England: implication for the origin and controls
of the Black Sea. Deep-Sea Research 38, S921–S942. of the distribution of mudstones. Journal of Sedimentary Research 63, 1129–1143.
Karl, D., Michaels, A., Bergman, B., Capone, D., Carpenter, E., Letelier, R., Lipschultz, F., Manheim, F.T., Pratt, R.M., McFarlin, 1980. Composition and origin of phosphate deposits
Paerl, H., Sigman, D., Stal, L., 2002. Dinitrogen fixation in the world's oceans. of the Blake Plateau. In: Bentor, Y.K. (Ed.), Marine Phosphorites. Society of Economic
Biogeochemistry 57, 47–98. Paleontologists and Mineralogists, Special Publication, vol. 29, pp. 117–137.
Karstensen, J., Stramma, L., Visbeck, M., 2008. Oxygen minimum zones in the eastern Mann, K.H., Lazier, J.R.N., 2006. Dynamics of Marine Ecosystems: Biological–physical
tropical Atlantic and Pacific oceans. Progress in Oceanography, 77, 331–350. Interactions in the Oceans. Blackwell Scientific Publications, Boston, MA. 496 pp.
Kennett, J.P., Watkins, N.D., 1975. Deep-sea erosion and manganese nodule develop- Marcantonio, F., Anderson, R.F., Higgins, S.M., Stute, M., Schlosser, P., Kubik, P., 2001. Sediment
ment in the southeast Indian Ocean. Science 188, 1011–1013. focusing in the central equatorial Pacific Ocean. Paleoceanography 16, 260–267.
Kienast, S.S., Kienast, M., Mix, A.C., Calvert, S.E., Francois, R., 2007. Thorium-230 Martin, J.H., 1991. Iron, Liebig's Law and the greenhouse. Oceanography 4, 52–55.
normalized particle flux and sediment focusing in the Panama Basin region during Martin, J.H., 1992. Iron as a limiting factor in oceanic productivity. In: Falkowski, P.G.,
the last 30,000 years. Paleoceanography 22, PA2213. doi:10.1029/2006PA001357. Woodhead, A.D. (Eds.), Primary Productivity and Biogeochemical Cycles in the Sea.
Kim, K.H., Burnett, W.C., 1988. Accumulation and biological mixing of Peru margin Environmental Science Research, vol. 43. Plenum Press, New York, pp. 123–137.
sediments. Marine Geology 80, 181–194. Martin, J.H., Knauer, G.A., 1973. The elemental composition of plankton. Geochimica et
King, C., 2004. The Black Sea: A History. Oxford University Press, Oxford. 276 pp. Cosmochimica Acta 37, 1639–1651.
Knox, F., McElroy, M.B., 1984. Changes in atmospheric CO2: influence of the marine biota Martin, J.M., Meybeck, M., 1979. Elemental mass-balance of material carried by major
at high latitude. Journal of Geophysical Research 89, 4629–4637. world rivers. Marine Chemistry 7, 173–206.
Komar, P.D., Neudeck, R.H., Kulm, L.D., 1972. Observation and significance of deep-water Martinez, J.I., 2003. The paleoecology of Late Cretaceous upwelling events from the
oscillatory ripple marks on the Oregon continental shelf. In: Swift, D.J.P., Duane, D.B., Upper Magdalena Basin, Colombia. Palaios, 18, 305–320.
Pilkey, O.H. (Eds.), Shelf Sediment Transport: Process and Pattern. Dowden, Maybeck, M., 1982. Carbon, nitrogen and phosphorus transport by world rivers.
Hutchinson and Ross, Stroudsburg, Pennsylvania, pp. 601–619. American Journal of Science 282, 401–450.
Kremling, K., 1983. The behavior of Zn, Cd, Cu, Ni, Co, Fe and Mn in anoxic Baltic waters. McCartney, M.S., 1977. Subantarctic mode water. In: Angel, M.V. (Ed.), A Voyage of
Marine Chemistry 13, 87–108. Discovery. Pergamon, Oxford, pp. 103–119.
Kuehl, S.A., Nittouer, C.A., Allison, M.A., Ercilio, L., Faria, C., Dukat, D.A., Jaeger, J.M., McElroy, M.B., 1983. Marine biological controls on atmospheric CO2 and climate. Nature
Pacioni, T.D., Figueiredo, A.G., Underkoffler, E.C., 1996. Sediment deposition, 302, 328–329.
accumulation, and seabed dynamics in an energetic fine-grained coastal environment. McManus, J., Berelson, W.B., Severmann, S., Poulson, R.L., Hammond, D.E., Klikhammer,
Continental Shelf Research 16, 787–815. G.P., Holm, C., 2006. Molybdenum and uranium geochemistry in continental margin
Kuypers, M.M.M., Lavik, G., Woebken, D., Schmid, M., Fuchs, B.M., Amann, R., Jorgensen, sediments: paleoproxy potential. Geochimica et Cosmochimica Acta 70, 4643–4662.
B.B., Jetten, M.S.M., 2005. Massive nitrogen loss from the Benguela upwelling Miller, R.G., 1990. A paleoceanographic approach to the Kimmeridge Clay Formation. In:
system through anaerobic ammonium oxidation. Proceedings of the National Huc, A.Y. (Ed.), Deposition of Organic Facies. American Association of Petroleum
Academy of Sciences 102, 6478–6483. Geologists Studies in Geology, vol. 30, pp. 13–26. Tulsa, OK.
Kuypers, M.M.M., Sleiders, A.O., Lavik, G., Schmid, M., Jorgensen, B.B., Kuenen, J.G., Miller, K.G., Kominz, M.A., Browning, J.V., Wright, J.D., Mountain, G.S., Katz, M.E.,
Sinninghe Damste, J.S., Strous, M., Jetten, M.S.M., 2003. Anaerobic ammonium Sugarman, P.J., Cramer, B.S., Christe-Blick, N., Pekar, S.F., 2005. The Phanerozoic
oxidation by anammox bacteria in the Black Sea. Nature 422, 608–610. record of global sea-level change. Science 310, 1293–1298.
Lallier-Vergès, E., Hayes, J.M., Boussafir, M., Zabeck, D.A., Tribovillard, N.P., Connan, J., Millero, F.J., 2001. The Physical Chemistry of Natural Water. J. Wiley, New York.
Bertrand, P., 1997. Productivity-induced sulphur enrichment of hydrocarbon- Milliman, J.D., Boyle, E., 1975. Biological uptake of dissolved silica in the Amazon River
rich sediments from the Kimmeridge Clay Formation. Chemical Geology 134, Estuary. Science 189, 995–997.
277–288. Moore, W.S., Sarmiento, J.L., Key, R.M., 1986. Tracing the Amazon component of surface
Lallier-Vergès, E., Tribovillard, N.-P., Bertrand, P., 1995. Organic Matter Accumulation. Atlantic water using 228Ra, salinity and silica. Journal of Geophysical Research 91,
Springer-Verlag, Berlin, 187 pp. 2574–2580.
94 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

Morford, J.L., Emerson, S., Breckel, E.J., Kim, S.H., 2005. Diagenesis of oxyanions (V, U, Re, Paerl, H.W., 1993. Emerging role of atmospheric nitrogen deposition in coastal
and Mo) in pore waters and sediments from a continental margin. Geochimica et eutrophication: biogeochemical and tropic perspectives. Canadian Journal of
Cosmochimica Acta 21, 5021–5032. Fisheries and Aquatic Sciences 50, 2254–2269.
Morgans-Bell, H.S., Coe, A., Hesselbo, S.P., Jenkyns, H.C., Weedon, G.P., Marshall, J.E.A., Paerl, H., 1995. Coastal eutrophication in relation to atmospheric nitrogen deposition:
Williams, C.J., 2001. The stratotype of the Kimmeridge Clay Formation (Upper current perspectives. Ophelia 41, 237–259.
Jurassic) based on exposures and boreholes in south Dorset, U.K. Geological Parrish, J.T., 1987. Palaeo-upwelling and the distribution of organic-rich rocks. In:
Magazine 138, 511–539. Brooks, J., Fleet, A.J. (Eds.), Marine Petroleum Source Rocks. The Geological Society
Morrison, J., Smith, O.P., 1990. Geostrophic transport variability along Aves Ridge in the (London), Special Publication, vol. 26, pp. 199–205.
eastern Caribbean Sea during 1985–1986. Journal of Geophysical Research 95, 699–710. Pearson, S.J., Marshall, J.E.A., Kemp, A.E.S., 2004. The White Stone Band of the
Morse, J.W., Emeis, K.C., 1990. Controls on C/S ratios in hemipelagic upwelling Kimmeridge Clay: an integrated high-resolution approach to understanding
sediments. American Journal of Science 290, 1117–1135. environmental change. Journal of the Geological Society 161, 675–684.
Moucha, R., Forte, A.M., Mitrovica, J.X., Rowley, D.B., Quéré, S., Simmons, N.A., Grand, S.P., Perissinotto, R., Laubscher, R.K., McQuaid, C.D., 1992. Marine productivity enhancement
2008. Dynamic topography and long-term sea-level variations: there is no such thing around Bouvet and the South Sandwich Islands (Southern Ocean). Marine Ecology
as a stable continental platform. Earth and Planetary Science Letters 271, 101–108. Progress Series 88, 41–53.
Müller, G., Blaschke, R., 1969. Zur Entstehung des Tiefsee-Kalkschlammes im Schwarzen Perkins, R.B., Piper, D.Z., Mason, C.E., 2008. Trace-element budgets in the Ohio-Sunbury
Meer. Naturwissenschaften 56, 561–562. shales of Kentucky: constraints on ocean circulation and primary productivity in the
Müller, G., Stoffers, P., 1974. Mineralogy and petrology of Black Sea basin sediments. In: Devonian–Mississippian Appalachian Basin. Palaeogeography Palaeoclimatology
Degens, E.T., Ross, D.A. (Eds.), The Black Sea—Geology, Chemistry and Biology. Palaeoecology 265, 14–29.
American Association of Petroleum Geologists Memoir, vol. 20, pp. 200–248. Tulsa, Peterson, L.C., Overpeck, J.T., Kipp, N.G., Imbrie, J., 1991. A high-resolution late Quaternary
OK. upwelling record from the anoxic Cariaco Basin, Venezuela. Paleoceanography 6,
Müller-Karger, F.E., McClain, C.R., Fisher, T.R., Esaias, W.E., Varela, R., 1989a. Pigment 99–120.
distribution in the Caribbean Sea: observations from space. Progress in Oceanography Peterson, L.C., Overpeck, J.T. and Murray, D.W., 1995. Anoxic basin records detailed
23, 23–64. climate history. Drill Bits (JOI/USSAC Newsletter) 8, 10–13.
Müller-Karger, F.E., McClain, C.R., Richardson, P.L., 1989b. The dispersal of the Amazon's Pilipchuk, M.F., Volkov, I.I., 1974. Behavior of molybdenum in processes of sediment
water. Nature 333, 56–59. formation and diagenesis in Black Sea. In: Degens, E.T., Ross, D.A. (Eds.), The Black Sea—
Müller-Karger, F., Varela, R., Walsh, J.J., 2001. Annual cycle of primary production in the Geology, Chemistry and Biology. American Association of Petroleum Geologists
Cariaco Basin: response to upwelling and implications for vertical export. Journal of Memoir, vol. 20, pp. 542–553. Tulsa, OK.
Geophysical Research 106, 4527–4542. Piper, D.Z., 1994. Seawater as the source of minor elements in black shales, phosphorites
Müller-Karger, F.E., Varela, R., Thunell, R., Luerssen, R., Hu, C., Walsh, J.J., 2005. The and other sedimentary rocks. Chemical Geology 114, 95–114.
importance of continental margins in the global carbon cycle. Geophysical Research Piper, D.Z., 2001. Marine chemistry of the Permian Phosphoria Formation and Basin,
Letters 32, L01602. doi:10.1029/2004GL021346. Southeast Idaho. Economic Geology 96, 599–620.
Mullin, M.M., 1986. Spatial and temporal patterns. In: Eppley, R.W. (Ed.), Plankton Piper, D.Z. and Dean, W.E., 2002. Trace-element deposition in the Cariaco Basin, Venezuela
Dynamics of the Southern California Bight. Lecture Notes on Coastal and Estrarine Shelf, under sulfate-reducing conditions—a history of the local hydrography and global
Studies, vol. 15. Springer, Berlin, pp. 216–273. climate, 20 ka to the present. Professional Paper 1670, U.S. Geological Survey,
Murray, J.W., 1991. The 1988 Black Sea oceanographic expedition, introduction and Washington, D.C., 41 pp; http://geopubs.wr.usgs.gov/prof-paper/pp1670/.
summary. Deep-Sea Research 38, S655–S661. Piper, D.Z. and Isaacs, C.M., 1994. Geochemistry of minor elements in the Monterey
Murray, J.W., Jannasch, H.W., Honjo, S., Anderson, R.F., Reeburgh, W.S., Top, Z., Formation, CA: Seawater chemistry of deposition, U. S. Geological Survey Professional
Friederich, G.E., Codispoti, L.A., Izdar, E., 1989. Unexpected changes in the oxic/ Paper 1566, Reston, VA., 44 pp.
anoxic interface in the Black Sea. Nature 338, 411–413. Piper, D.Z., Isaacs, C.M., 1996. Instability of bottom-water redox conditions during
Murray, J.W., Spell, B., Paul, B., 1983. The contrasting geochemistry of manganese and accumulation of Quaternary sediment in the Japan Sea. Paleoceanography 11, 171–190.
chromium in the eastern tropical Pacific Ocean. In: Wong, C.S., Boyle, E., Bruland, K.W., Piper, D.Z., Wandless, G.A., 1992. Hydroxylamine-hydrochloride-acetic-acid-soluble-
Goldberg, E.D. (Eds.), Trace Metals in Seawater, NATO Conference Series IV. Plenum and insoluble-fractions of pelagic sediment—readsorption revisited. Environmental
Press, New York, pp. 643–669. Science and Technology 26, 2489–2493.
Murray, J.W., Stewart, K., Kassakian, S., Krynytzky, M., DiJulio, D., 2007. Oxic, suboxic, and Piper, D.Z., Perkins, R.B., Rowe, H.D., 2007. Rare-earth elements in the Permian
anoxic conditions in the Black Sea. In: Yanko-Hombach, V., Gilbert, A.S., Panin, N., Phosphoria Formation: paleo proxies of ocean geochemistry. Deep-Sea Research 54,
Dolukhaov, P.M. (Eds.), The Black Sea Flood Question: Changes in Coastline, Climate, 1396–1413.
and Human Settlement. Springer, Dordrecht, pp. 1–21. Piper, D.Z., Rude, P.D., Monteith, S., 1986. Chemistry, mineralogy, and age of haloed
Murray, J.W., Top, Z., Özsoy, E., 1991. Hydrographic properties and ventilation of the burrows in pelagic sediment: DOMES Site A in the equatorial North Pacific. Marine
Black Sea. Deep-Sea Research 38, S663–S689. Geology 74, 41–55.
Nameroff, T.J., Balistrieri, L.S., Murray, J.W., 2002. Suboxic trace metal geochemistry in the Pitcher, G.C., Brown, P.C., Mitchell-Innes, B.A., 1992. Spatio-temporal variability of
Eastern Tropical North Pacific. Geochimica et Cosmochimica Acta 66, 1139–1158. phytoplankton in the southern Benguela Upwelling system. South African Journal
Nelson, G., Hutchings, L., 1983. The Benguela upwelling area. Progress in Oceanography 12, of Marine Science, 12, 439–456.
333–356. Polat, S.C., Tuğrul, S., 1995. Nutrient and organic carbon exchanges between the Black
Nirel, P.M.V., Morel, F.M.M., 1990. Pitfalls of sequential extractions. Water Research 24, and Marmara Seas through the Bosphorus Strait. Continental Shelf Research 15,
1055–1056. 1115–1132.
Nixon, S.W., Ammerman, L.P., Atkinson, V.M., Berounsky, G.D., Elmgren, R., Garber, J.H., Proust, J.N., Deconinck, J.F., Geyssant, J.R., Herbin, J.P., Vidier, J.P., 1995. Sequence
Giblin, A.E., Janke, R.A., Owens, N.J.P., Pilson, M.E.Q., Seitzinger, S.P., 1996. The fate of analytical approach to the Upper Kimmeridgian–Lower Tithonian storm-domi-
nitrogen and phosphorus at the land-sea margin of the North Atlantic Ocean. nated ramp deposits of the Boulonnais (Northern France). A landward time-
Biogeochemistry 35, 141–180. equivalent to offshore marine source rocks. Geologische Rundschau 84, 255–271.
Nolan, B.T., Stoner, J.D., 2000. Nutrients in groundwaters of the conterminous United Quigg, A., Finkel, Z.V., Irwin, A.J., Rosenthal, Y., Ho, T.-Y., Reinfelder, J.R., Schofield, O.,
States, 1992–1995. Environmental Science and Technology 34, 1156–1165. Morel, M.M., Falkowski, P.G., 2003. The evolutionary inheritance of elemental
Oguz, T., 2005. Long-term impacts of anthropogenic forcing on the Black Sea ecosystem. stoichiometry in marine phytoplankton. Nature 425, 291–293.
Oceanography and Marine Biology Annual Reviews 18, 112–121. Rabalais, N.N., Turner, R.E., Justic, D., Dortch, Q., Wiseman, W.J., Sen Gupta, B.K., 1996.
Oschlies, A., 2002. Nutrient supply to surface waters of the North Atlantic: a model Nutrient changes in the Mississippi River and system responses on the adjacent
study. Journal of Geophysical Research 107. doi:10.1029/2000JC000275. continental shelf. Estuaries 19, 386–407.
Oschmann, W., 1988a. Upper Kimmeridgian and Portlandian marine macrobenthic Raiswell, R., Berner, R.A., 1985. Pyrite formation in euxinic and semi-euxinic sediments.
associations from southern England and northern France. Facies 18, 49–82. American Journal of Science 285, 710–724.
Oschmann, W., 1988b. Kimmeridge Clay sedimentation—a new cyclic model. Palaeo- Raiswell, R., Newton, R., Wignall, P.B., 2001. An indicator of water-column anoxia:
geography Palaeoclimatology Palaeoecology 65, 217–251. resolution of biofacies variations in the Kimmeridge Clay (Upper Jurassic, U.K.).
Oschmann, O., 1990. Environmental cycles in the late Jurassic northwest European epeiric Journal of Sedimentary Research 71, 286–294.
basin: interaction with atmospheric and hydrospheric circulations. Sedimentary Ramanampisoa, L., Disnar, J.R., 1994. Primary control of paleoproduction on organic
Geology 69, 313–323. matter preservation and accumulation in the Kimmeridge rocks of Yorkshire (UK).
Ostlund, H.G., 1974. Expedition “Odysseus 65” radiocarbon age of Black Sea deep water. In: Organic Geochemistry 21, 1153–1167.
Degens, E.T., Ross, R.A. (Eds.), The Black Sea—Geology, Chemistry, and Biology. Redfield, A.C., Ketchum, B.H., Richards, F.A., 1963. The influence of organisms on the
American Association of Petroleum Geologists Memoir, vol. 20, pp. 127–132. Tulsa, OK. composition of sea water. In: Hill, M.N. (Ed.), The Sea, vol. 2. Wiley, New York, pp. 26–77.
Özsoy, E., Ünlüata, Ü., 1997. Oceanography of the Black Sea: a review of some recent Rex, R.W., Goldberg, E.D., 1958. Quartz contents of pelagic sediments in the Pacific
results. Earth-Science Reviews 42, 231–272. Ocean. Tellus 10, 153–159.
Özsoy, E., DiIorio, D., Gregg, M.C., Backhaus, J.O., 2001. Mixing in the Bosphorus Strait Richards, F.A., 1975. The Cariaco Basin (Trench). Oceanography and Marine Biology
and the Black Sea continental shelf: observations and a model of the dense water Annual Reviews 13, 11–67.
outflow. Journal of Marine Systems 31, 99–135. Rimmer, S.M., Thompson, J.A., Goodnight, S.A., Robl, T.L., 2004. Multiple controls on the
Özsoy, E., Top, Z., White, G.J., Murray, J.W., 1991. Double diffusive intrusions, mixing and preservation of organic matter in Devonian–Mississippian marine black shales:
deep sea convection processes in the Black Sea. In: Izdar, E., Murray, J.W. (Eds.), Black geochemical and petrographic evidence. Palaeogeography Palaeoclimatology
Sea Oceanography. Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 17–42. Palaeoecology 215, 125–154.
NATO ASI Series. Rutkowski, C.M., Burnett, W.C., Iverson, R.L., Chanton, J.P., 1999. The effect of
Pace, M.G., Knauer, D.K., Martin, J., 1987. Primary production, new production and groundwater seepage on nutrient delivery and seagrass distribution in the
vertical flux in the eastern Pacific Ocean. Nature 325, 803–804. northeastern Gulf of Mexico. Estuaries 22, 1033–1040.
D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96 95

Ryabchenko, V., Gorchakov, V., Fasham, M., 1998. Seasonal dynamics and biological Suess, E., 1980. Particulate organic carbon flux in the ocean—surface productivity and
productivity in the Arabian Sea euphotic zone as simulated by a three-dimensional oxygen utilization. Nature 288, 260–263.
ecosystem model. Global Biogeochemical Cycles 12, 501–530. Suits, N.S., Arthur, M.A., 2000. Sulfur diagenesis and partitioning in Holocene Peru shelf
Ryther, J.H., Menzel, D.W., Corwin, N., 1967. Influence of the Amazon River outflow on and upper slope sediments. Chemical Geology 163, 219–234.
the ecology of the western Tropical Atlantic I. Hydrography and nutrient chemistry. Suzuki, N., Ishida, K., Shinomiya, Y., Ishiga, H., 1998. High productivity in the earliest
Journal of Marine Research 25, 69–83. Triassic ocean: black shales, Southwest Japan. Palaeogeography Palaeoclimatology
Sadiq, M., 1988. Thermodynamic solubility relationships of inorganic vanadium in the Palaeoecology 141, 53–65.
marine environment. Marine Chemistry 23, 87–96. Sverdrup, H.U., Johnson, M.W., Fleming, R.H., 1942. The Oceans. Their Physics,
Sahagian, D., Pinous, O., Olferiev, A., Zakharov, V., 1996. Eustatic curve for the Middle Chemistry, and General Biology. Prentice-Hall, Englewood Cliffs, NJ. 1087 pp.
Jurassic–Cretaceous based on Russian Platform and Siberian stratigraphy; zonal Taylor, A.H., Harbour, D.S., Harris, R.P., Burkill, P.H., Edwards, E.S., 1993. Seasonal
resolution. American Association of Petroleum Geologists Bulletin 80, 1433–1458. succession in the pelagic ecosystem of the North Atlantic and the utilization of
Saelen, G., Tyson, R.V., Telnaes, N., Talbot, M.R., 2000. Contrasting watermass conditions nitrogen. Journal of Plankton Research 15, 875–891.
during deposition of the Whitby Mudstone (Lower Jurassic) and Kimmeridge Clay Taylor, S.R., McLennan, S.M., 1985. The Continental Crust, its Composition and Evolution.
(Upper Jurassic) Formations. Palaeogeography Palaeoclimatology Palaeoecology Blackwell Scientific, Oxford. 312 pp.
163, 163–196. Tessier, A., Campbell, P.G.C., 1991. Comments on Pitfalls of sequential extractions by P.M.V.
Sarmiento, J.L., Gruber, N., 2006. Ocean Biogeochemical Dynamics. Princeton University Nirel and F.M.M. Morel. Water Research 25, 115–117.
Press, Princeton. 503 pp. Tessier, A., Campbell, P.G.C., Bisson, M., 1979. Sequential extraction procedure for the
Sarmiento, J.L., Gruber, N., Brzezinski, M.A., Dunne, J.P., 2004. High-latitude controls of speciation of particulate trace metals. Analytical Chemistry 51, 844–851.
thermocline nutrients and low latitude biological productivity. Nature 427, 56–60. Thunell, R.C., Varela, R., Bohrer, R., 2000. Organic carbon fluxes, degradation, and
Savrda, C.E., Bottjer, D.J., 1987. The exaerobic zone, a new oxygen-deficient marine accumulation in an anoxic basin: sediment trap results from the Cariaco Basin.
biofacies. Nature 327, 54–56. Limnology and Oceanography 45, 300–308.
Schmidt, I., Sliekers, O., Schmid, M., Cirpus, I., Strous, M., Bock, E., Kuenen, J.G., Jetten, M.S.M., Toggweiler, J.R., Russell, J.L., Carson, S.R., 2006. Midlatitude westerlies atmospheric CO2
2002. Aerobic and anaerobic ammonia oxidizing bacteria—competitors or natural and climate change during the ice ages. Paleoceanography 21, PA2005. doi:10.1029/
partners? FEMS Microbiology Ecology 39, 175–181. 2005PA001154.
Sclater, F.R., Boyle, E., Edmond, J.M., 1976. On the marine geochemistry of nickel. Earth Tourtelot, H.A., 1979. Black shale: its deposition and diagenesis. Clays and Clay Minerals
and Planetary Science Letters 31, 119–128. 27, 313–321.
Scotese, C.R., 2001. Paleomap project. http://www.scotese.com/earth.htm. Trask, P.D., 1932. Origin and Environment of Source Sediments of Petroleum. Gulf
Scranton, M.I., Astor, Y., Bohrer, R., Ho, T.-Y., Müller-Karger, F., 2001. Controls on temporal Publishing Company, Houston, TX. 323 pp.
variability of the geochemistry of the deep Cariaco Basin. Deep Sea Research II 48, Tribovillard, N., Bialkkowski, A., Tyson, R.V., Lallier-Verges, E., Deconinck, 2001. Organic
1605–1625. facies variation in the late Kimmeridgian of the Boulonnais area (northernmost
Shepard, F.P., Phleger, F.B., van Andel, T.H. (Eds.), 1960. Recent Sediments, Northwest France). Marine and Petroleum Geology 18, 371–389.
Gulf of Mexico. American Association of Petroleum Geologists. Tulsa, OK, 394 pp. Tribovillard, N., Desprairies, A., Lallier-Vèrges, E., Bertrand, P., Moureau, N., Ramdani, A.,
Sholkovitz, E.R., 1989. Artifacts associated with the chemical leaching of sediments for Ramanampisoa, L., 1994. Geochemical study of organic-matter rich cycles from the
rare-earth elements. Chemical Geology 77, 47–51. Kimmeridge Clay Formation of Yorkshire (UK): productivity versus anoxia.
Sholkovitz, E.R., 1993. Comment on “Redox cycling of rare earth elements in the suboxic Palaeogeography Palaeoclimatology Palaeoecology 108, 165–181.
zone of the Black Sea” by C.R. German, B.P. Holliday and H. Elderfield, 1991. Tribovillard, N., Lyons, T.W., Riboulleau, A., Bout-Roumazeilles, 2008. A possible capture
Geochimica et Cosmochimica Acta 56, 4305–4307. of molybdenum during early diagenesis of dysoxic sediments. Société Géologique
Showers, W.J., Angle, D.G., 1986. Stable isotopic characterization of organic carbon de France Bulletin 179, 3–12.
accumulation on the Amazon continental shelf. Continental Shelf Research 6, 227–244. Tribovillard, N., Ramdani, A., Trentesaux, A., 2005. Controls on organic accumulation in
Sillén, L.G., 1961. The physical chemistry of seawater. In: Sears, M. (Ed.), Oceanography. Upper Jurassic shales of northwestern Europe as inferred from trace-metal
American Association for the Advancement of Science Publication, vol. 67, pp. 549–581. geochemistry. In: Harris, N.B. (Ed.), The Deposition of Organic-carbon-rich Sediments:
Washington, D.C. Models, Mechanisms, and Consequences. Special Publication, vol. 82. Society for
Sinninghe Damsté, J.S., de Leeuw, J.W., 1990. Analysis, structure and geochemical Sedimentary Geology, Tulsa, OK, pp. 145–164.
significance of organically-bound sulphur in the geosphere: state of the art and Turekian, K.K., Wedepohl, L.H., 1961. Distribution of the elements in some major units of
future research. In: Durand, B., Behar, F. (Eds.), Advances in Organic Geochemistry. the Earth's crust. Geological Society of America Bulletin 72, 175–192.
Proceedings of the 14th International Meeting on Organic Geochemistry. Pergamon Turner, R.J.W., 1992. Formation of Phanerozoic stratiform sediment-hosted zinc–lead
Press, Oxford, pp. 1077–1101. deposits: evidence for the critical role of ocean anoxia. Chemical Geology 99, 165–188.
Slomp, C.P., Van Cappellen, P., 2004. Nutrient inputs to the coastal ocean through Tyrrell, T., Lucas, M.I., 2002. Geochemical evidence of denitrification in the Benguela
submarine groundwater discharge: controls and potential impact. Journal of upwelling system. Continental Shelf Research 22, 2497–2511.
Hydrology 295, 64–86. Tyson, R.V., 1987. The genesis and palynofacies characteristics of marine petroleum
Smith, R.L., 1995. The physical processes of coastal ocean upwelling systems. In: source rocks. In: Brooks, J., Fleet, A.J. (Eds.), Marine Petroleum Source Rocks.
Summerhayes, C.P., Emeis, K.-C., Angel, M.V., Smith, R.L., Zeitzschel, B. (Eds.), Geological Society Special Publication, vol. 26, pp. 47–67. London.
Upwelling in the Ocean: Modern Processes and Ancient Records. Environmental Tyson, R.V., 1995. Sedimentary Organic Matter: Organic Facies and Palynofacies.
Sciences Research Report ES-18. J. Wiley, New York, pp. 39–64. Chapman and Hall, London. 615 pp.
Smith, W.O., DeMaster, D.J., 1996. Phytoplankton biomass and productivity in the Tyson, R.V., 1996. Sequence-stratigraphical interpretation of organic facies variations in
Amazon River plume: correlation with seasonal river discharge. Continental Shelf marine siliclastic systems: general principles and application to the onshore
Research 16, 291–319. Kimmeridge Clay formation, UK. In: Hesselbo, S.P., Parkinson, D.N. (Eds.), Sequence
Sommerfield, C.K., Nittrouer, C.A., DeMaster, D.J., 1996. Sedimentary carbon-isotope Stratigraphy in British Geology. Geological Society Special Publication, vol. 103.
systematics on the Amazon shelf. Geo-Marine Letters 16, 17–23. London, pp. 75–96.
Sorokin, Y.I., 1983. The Black Sea. In: Ketchum, B.H. (Ed.), Ecosystems of the World 26. Tyson, R.V., 2004. Variation in total organic carbon through the type Kimmeridge Clay
Elsevier, Amsterdam, pp. 253–292. Formation (Late Jurassic), Dorset, UK. Journal of the Geological Society (London)
Sorokin, Y.I., 1999. Data on primary production in the Bering Sea and adjacent Northern 161, 667–673.
Pacific. Journal of Plankton Research 21, 615–636. Tyson, R.V., 2005. The “productivity versus preservation” controversy: cause, flaws and
Soutar, A., Crill, P., 1977. Sedimentation and climatic patterns in the Santa Barbara Basin resolution. In: Harris, N.B. (Ed.), The Deposition of Organic-carbon-rich Sediments:
during the 19th and 20th centuries. Geological Society of America Bulletin 88, Models, Mechanisms, and Consequences. Special Publication, vol. 82. Society for
1161–1172. Sedimentary Geology, Tulsa, OK. 17–33.
Southard, J.B., Cacchoine, D.A., 1972. Experiments on bottom sediment movement by Tyson, R.V., Pearson, T.H., 1991. Modern and ancient continental shelf anoxia: an
breaking internal waves. In: Swift, D.J.P., Duane, D.B., Pilkey, O.H. (Eds.), Shelf overview. In: Tyson, R.V., Pearson, T.H. (Eds.), Modern and Ancient Continental
Sediment Transport: Process and Pattern. Dowden, Hutchinson and Ross, Strouds- Shelf Anoxia. Geological Society Special Publications, vol. 58. The Geological Society
burg, Pennsylvania, pp. 83–97. (London), pp. 1–24.
Spears, D.A., Kanaris-Sotiriou, R., 1976. Titanium in some Carboniferous sediments from Tyson, R.V., Wilson, R.C.L., Downie, C., 1979. A stratified water column environmental
Great Britain. Geochimica et Cosmochimica Acta 40, 345–351. model for the type Kimmeridge Clay. Nature 277, 377–380.
Stanev, E.V., 2005. Understanding Black Sea dynamics. Oceanography and Marine van Andel, T.H., Calvert, S.E., 1971. Evolution of a sediment wedge, Walvis shelf,
Biology Annual Reviews 18, 56–75. southwest Africa. Journal of Geology 79, 585–602.
Stenseth, N.C., Ottersen, G., Hurrell, J.W., Belgrano, A. (Eds.), 2004. Marine ecosystems van Andel, T.H., Shor, G.G. (Eds.), 1964. Marine Geology of the Gulf of California.
and climate variation. The North Atlantic: A Comparative Perspective. Oxford American Association of Petroleum Geologists, Tulsa, OK. 408 pp.
University Press, Oxford. 252 pp. van Bennekom, A.J., Berger, G.W., Helder, W., Vries, R.T.P.D., 1978. Nutrient distribution in
Sternberg, R.W., Cacchione, D.A., Paulson, B., Kineke, G.C., Drake, D.E., 1996. Observations of the Zaire estuary and river plume. Netherlands Journal of Sea Research 12, 296–323.
sediment transport on the Amazon subaqueous delta. Continental Shelf Research 16, van Bennekom, A.J., Gieskes, W.W.C., Tijssen, S.B., 1975. Eutrophication of Dutch coastal
697–715. waters. Proceedings of the Royal Society (London) 189B, 359–374.
Strakhov, N.M., 1971. Geochemical evolution of the Black Sea in the Holocene. Lithology van Dongen, B.E., Schouten, S., SinningheDamste, J.S., 2006. Preservation of carbohydrates
and Mineral Resources 3, 263–274. through sulfurization in a Jurassic euxinic shelf sea: examination of the Blackstone
Stow, D.A.V., Huc, A.-Y., Bertrand, P., 2001. Depositional processes of black shales in deep Band TOC cycle in the Kimmeridge Clay Formation, UK. Organic Geochemistry 37,
water. Marine and Petroleum Geology 18, 491–498. 1052–1073.
Stumm, W., Morgan, J.J., 1996. Aquatic Chemistry: Chemical Equilibria and Rates in van Dover, C.L., 2000. The Ecology of Deep-sea Hydrothermal Vents. Princeton
Natural Waters. J. Wiley, New York. University Press, Princeton, N.J.. 424 pp.
96 D.Z. Piper, S.E. Calvert / Earth-Science Reviews 95 (2009) 63–96

Varela, R., Capelo, J.C., Gutierrez, J., Müller-Karger, F., Diaz-Ramos, J.R., 1997. Primary Wignall, P.B., 1989. Sedimentary dynamics of the Kimmeridge Clay: tempests and
productivity in Cariaco Basin waters. Transactions of the American Geophysical earthquakes. Journal of the Geological Society (London) 146, 273–284.
Union 78, F342. Wignall, P.B., 1990. Benthic palaeoecology of the Late Jurassic Kimmeridge clay of England.
Varela, I., Costa, J., Foreman, K., Teal, J.M., Howes, B.L., Aubrey, D., 1990. Transport of The Palaeontological Association, Special Papers in Palaeontology 43, 1–74.
groundwater-borne nutrients from watersheds and their effects on coastal waters. Wignall, P.B., 1991. Dysaerobic trace fossils and ichnofabrics in the upper Jurassic
Biogeochemistry 10, 177–197. Kimmeridge Clay of southern England. Palaios 6, 264–270.
Vine, J.D., Tourtelot, E.B., 1970. Geochemistry of black shale deposits—a summary report. Wignall, P.B., 1994. Black Shales. Clarendon Press, Oxford. 127 pp.
Economic Geology 65, 253–272. Wignall, P.B., Hallam, A., 1991. Biofacies, stratigraphic distribution and depositional
Volkov, I.I., Fomina, L.S., 1974. Influence of organic material and processes of sulfide models of British onshore Jurassic black shales. In: Tyson, R.V., Pearson, T.H. (Eds.),
formation on distribution of some trace elements in deep-water sediments of the Modern and Ancient Continental Shelf Anoxia. Geological Society Special Publica-
Black Sea. In: Degens, E.T., Ross, D.A. (Eds.), The Black Sea—Geology, Chemistry, and tions, vol. 58. The Geological Society (London), pp. 291–309.
Biology. American Association of Petroleum Geologists Memoir, vol. 20, pp. 456–476. Wignall, P.B., Myers, K.J., 1988. Interpreting benthic oxygen levels in mudrocks: a new
Tulsa, OK. approach. Geology 16, 452–455.
Von Damm, K.L., 1990. Seafloor hydrothermal activity: Black smoker chemistry and Wignall, P.B., Newton, R., 2001. Black shales on the basin margin: a model based on
chimneys. Annual Review of Earth and Planetary Sciences 18, 173–204. examples from the Upper Jurassic of the Boulonnais, northern France. Sedimentary
Wakeham, S.G., Ertel, J.R., 1988. Diagenesis of organic matter in suspended particles and Geology 144, 335–356.
sediments in the Cariaco Trench. Organic Geochemistry 13, 815–822. Wignall, P.B., Sutcliffe, O.E., Clemson, J., Young, E., 1996. Unusual shoreface sedimentol-
Walsh, J.J., 1981. Shelf-sea ecosystems. In: Alonghurst, A.R. (Ed.), Analysis of Marine ogy in the Upper Jurassic of the Boulonnais, northern France. Journal of
Ecosystems. Academic Press, London, pp. 159–196. Sedimentary Research 66, 577–586.
Walsh, J.J., 1991. Importance of continental margins in the marine biogeochemical Wijsman, J.W.M., Herman, P.M.J., Middelburg, J.J., Soetaert, K., 2002. A model for early
cycling of carbon and nitrogen. Nature 350, 53–55. diagenetic processes in sediments of the continental shelf of the Black Sea.
Watson, A.J., Naveira Garabato, A.C., 2006. The role of Southern Ocean mixing and Estuarine Coastal and Shelf Science 54, 403–422.
upwelling in glacial–interglacial atmospheric CO2 change. Tellus 58B, 73–87. Woolnough, W.G., 1937. Sedimentation in barred basins and source rocks of oil.
Wedepohl, K.H., 1971. Environmental influences on the chemical composition of shales American Association of Petroleum Geology Bulletin 21, 1101–1157.
and clays. In: Ahrens, L.H., Press, F., Runcorn, S.K., Urey, H.C. (Eds.), Physics and Wroblewski, J.S., Hofmann, E.E., 1989. U.S. interdisciplinary modeling studies of coastal–
Chemistry of the Earth. Pergamon, Oxford, pp. 307–331. offshore exchange processes: past and future. Progress in Oceanography 23, 65–99.
Wedepohl, K.H., 1995. The composition of the continental crust. Geochimica et Wyrtki, K., 1962. The oxygen minima in relation to ocean circulation. Deep-Sea Research
Cosmochimica Acta 59, 1217–1232. 9, 11–23.
Wedepohl, K.H., Correns, C.W., Shaw, D.M., Turekian, K.K., Zemann, J. (Eds.), 1969–1978. Wyrtki, K., 1963. The horizontal and vertical field of motion in the Peru Current. Scripps
Handbook of Geochemistry, I, II-1 through II-5. Springer, Berlin. Institution of Oceanography Bulletin 8, 313–346.
Weedon, G.P., Coe, A.L., Ogg, J.G., Gallois, R.W., 2004. Cyclostratigraphy, orbital tuning Yarincik, K.M., Murray, R.W., Peterson, L.C., 2000. Climatically sensitive eolian and
and inferred productivity for the type Kimmeridge Clay (Late Jurassic), southern hemipelagic deposition in the Cariaco Basin, Venezuela, over the past
England. Journal- Geological Society London 161, 655–666. 578,000 years: results from Al/Ti and K/Al. Paleoceanography 15, 210–228.
Wehrli, B., Stumm, W., 1989. Vanadyl in natural waters: adsorption and hydrolysis Yin, K., Harrison, P., Beamish, R.J., 1997. Effects of a fluctuation in Fraser River discharge
promote oxygenation. Geochimica et Cosmochimica Acta 53, 69–77. on primary production in the central Strait of Georgia, British Columbia, Canada.
Weiss, R.F., 1970. Solubility of nitrogen, oxygen, and argon in water and seawater. Deep- Canadian Journal of Fisheries and Aquatic Sciences 54, 1015–1024.
Sea Research 17, 721–735. Ziegler, P.A., 1990. Geological Atlas of Western and Central Europe. Shell International
Whitfield, M., 2001. Interactions between phytoplankton and trace metals in the ocean. Petroleum, Maatschappij B.V., Bath, England. 239 pp.
Advances in Marine Biology 41, 1–128. Zheng, Y., Anderson, R.F., Van Geen, A., Kuwabara, J., 2000. Authigenic molybdenum
Wiggert, J.D., Hood, R.R., Banse, K., Kindle, J.C., 2005. Monsoon-driven biogeochemical formation in marine sediments: a link to pore water sulfide in the Santa Barbara
processes in the Arabian Sea. Progress in Oceanography 65, 176–213. Basin. Geochimica et Cosmochimica Acta 64, 4165–4178.

You might also like