Journal Pre-proof Oceanic organic carbon as a possible first-order control on the carbon cycle during the Bathonian–Callovian
Ricardo L. Silva, Luís V. Duarte, Grant D. Wach, Natasha Morrison, Taylor Campbell PII:
S0921-8181(19)30543-0
DOI:
https://doi.org/10.1016/j.gloplacha.2019.103058
Reference:
GLOBAL 103058
To appear in:
Global and Planetary Change
Received date:
26 March 2019
Revised date:
5 October 2019
Accepted date:
14 October 2019
Please cite this article as: R.L. Silva, L.V. Duarte, G.D. Wach, et al., Oceanic organic carbon as a possible first-order control on the carbon cycle during the Bathonian–Callovian, Global and Planetary Change(2019), https://doi.org/10.1016/ j.gloplacha.2019.103058
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2019 Published by Elsevier.
Journal Pre-proof Oceanic organic carbon as a possible first-order control on the carbon cycle during the Bathonian–Callovian
Ricardo L. Silvaa,b,*
[email protected], Luís V. Duartec, Grant D. Wacha, Natasha Morrisona, Taylor Campbella
Basin and Reservoir Lab, Department of Earth Sciences, Dalhousie University, Life Sciences
of
a
iCRAG, Department of Geology, School of Natural Sciences, Trinity College Dublin, The
-p
b
ro
Center, 1355 Oxford Street, Halifax, Nova Scotia B3H 4R2, Canada
c
re
University of Dublin, College Green, Dublin 2, Ireland
MARE - University of Coimbra, Department of Earth Sciences, Rua Silvio Lima, Polo II, 3030-
na
*
lP
290, Coimbra, Portugal
Corresponding author at: Department of Geology, School of Natural Sciences, Trinity College
Jo
ur
Dublin, The University of Dublin, College Green, Dublin 2, Ireland.
Journal Pre-proof Abstract Oceans are the largest, readily exchangeable, superficial carbon reservoir; a current challenge in investigating past and present environments and predict future evolution relates to the role of oceanic carbon in regulating Earths’ carbon cycle and climate. At least one paired δ13Ccarb-TOC decoupling event is noted in the Late Bathonian–Early Callovian. Provokingly, we suggest that this decoupling event and other carbon isotopic events
of
in the Bathonian–Callovian resulted from the addition and removal of carbon from the oceanic
ro
organic carbon (OOC) reservoir, likely dominated by dissolved and fine particulate oceanic
-p
organic carbon. The decoupling event is characterised by a mainly invariant δ13CTOC record and
re
increasingly more positive δ13Ccarb values. The δ13Ccarb-TOC decoupling is tentatively explained by the expansion of the OOC reservoir, which increased the residence time of carbon in the oceans
lP
and buffered (slow equilibrium) δ13COOC (approximated by δ13CTOC) to changes in δ13C of
na
oceanic inorganic carbon, comprised mainly of dissolved inorganic carbon (approximated by δ13Ccarb). Reconversion of OOC to CO2 may have resulted in negative δ13C excursions and
ur
increased atmospheric pCO2, whereas the change from accumulation of OOC to export of
Jo
organic carbon to sediments may have resulted in positive δ13C excursions and widespread accumulation of organic matter. It is speculated that other small-scale (~ 1‰) δ13Ccarb-TOC excursions in the Mesozoic record may have resulted from similar episodes of OOC accumulation and oxidation. Our study calls attention to the potential of the OOC reservoir to act as a critical driver for planetary-scale changes over short geological time intervals. Keywords: δ13Ccarb-TOC; carbon cycle; palaeoceanography; Middle Jurassic; Lusitanian Basin; Portugal
Journal Pre-proof 1. Introduction The carbon cycle is usually divided into two cycles, the short- and long-term carbon cycles (Berner, 1999). Carbon in the short-term carbon cycle is rapidly exchanged within the surficial reservoirs, consisting of the oceans, atmosphere, biosphere, and soil, whereas carbon in the long-term cycle is slowly exchanged between the geosphere and the surficial systems (Berner, 1999; Ciais et al., 2013). The short-term carbon cycle is the dominant control on
of
atmospheric CO2 over millennia, whereas the long-term carbon cycle controls atmospheric CO2
ro
and O2 over millions of years (Berner, 1999, 2006).
-p
Oceans are, by far, the largest readily exchangeable superficial carbon reservoir; a current challenge in investigating past and present environments and predict future climatic evolution
re
concerns the role of the oceanic organic and inorganic carbon reservoirs in regulating Earths’
lP
carbon cycle and climate (Ciais et al., 2013; Hansell, 2013; Legendre et al., 2015). For example,
na
recent models suggest that ocean warming and freshening of surface waters is causing a slowdown of overturning circulation, which is projected to continue for the next millennia
ur
(Schmittner et al., 2008; Rahmstorf et al., 2015). It is postulated that accumulation of organic
Jo
carbon in the oceans might increase in these conditions, thus providing negative feedback to rising atmospheric CO2 (Jiao et al., 2010, 2014; Cavan et al., 2017; Shen and Benner, 2018).
1.1. Oceanic carbon in modern oceans Oceanic carbon is here divided into two large pools, oceanic inorganic carbon (OIC) and oceanic organic carbon (OOC), usually characterised as inorganic or organic or and then in dissolved or particulate (Fig. 1a). Dissolved inorganic carbon (DIC) consists of CO2* (=CO2 + H2CO3), HCO3−, CO32− in solution (Freeman and Hayes, 1992). Particulate inorganic carbon
Journal Pre-proof (PIC) results mainly from the precipitation of CaCO3 shells by coccolithophores, foraminifers, and pteropods, for example. On the other hand, OOC is operationally divided into dissolved organic carbon (DOC) and particulate organic carbon (POC), determined by the pore size of the filters (usually 0.7 μm or, less frequently 0.45 or 0.2 μm) used to separate “dissolved” from “particulate” material, respectively (Emerson and Hedges, 2008; Ciais et al., 2013; Carlson and Hansell, 2015) (Fig. 1b).
of
With a residence time of about 100 kyr (Emerson and Hedges, 2008), DIC is the largest
ro
carbon reservoir in today’s oceans (~38,000 PgC) (Ciais et al., 2013). The marine biota (~3 PgC)
-p
represents only a very small fraction of the oceanic carbon inventory and have a short residence
re
time (Ciais et al., 2013). In today’s oceans, recalcitrant DOC is the largest exchangeable carbon component of the OOC reservoir (Fig 1). The refractory DOC subclass holds about 630±32 PgC,
lP
similar to the size of the atmospheric carbon pool, consisting of ~832 PgC of which CO2 is the
na
dominant carbon-bearing gas and includes CH4 (~3.7 PgC) and CO (~0.2 PgC) (Ciais et al., 2013; Hansell, 2013). Refractory DOC comprises about 95% of the recalcitrant DOC reservoir
ur
and has an estimated residence time of about 16 kyr, indicating that oceans can store organic
Jo
carbon for extended periods of time and that geological variations in the size of the OOC reservoir, and particularly recalcitrant DOC, could impact atmospheric CO2 concentrations (Hansell, 2013) and, therefore, climate evolution and biological and geological processes. In this geological study, dissolved and particulate oceanic organic carbon (DPOOC) refers to organic carbon that is truly dissolved as well as suspended colloidal organic carbon (i.e. labile and recalcitrant DOC) and fine-grained POC (see discussions in (Ogawa and Tanoue, 2003; Verdugo et al., 2004; Fike et al., 2006; Emerson and Hedges, 2008; Swanson-Hysell et al.,
Journal Pre-proof 2010; Nebbioso and Piccolo, 2013) (Fig. 1b). The DPOOC component of the OOC reservoir is most likely dominated by recalcitrant DOC. Equilibrium and kinetic fractionation among organic matter (OM), CO2, and HCO3control the carbon isotope ratios of DIC, atmospheric CO2 and OOC (Emerson and Hedges, 2008). The main processes that affect the 13C/12C ratio of the “superficial” carbon reservoirs at geological time intervals are the equilibrium between DIC and atmospheric CO2, fractionation
of
between DIC and carbonate minerals; fixation of dissolved CO2 and production of biomass via
ro
photosynthesis, respiration, remineralisation, preservation, and sedimentary reworking and
-p
resuspension of OM in the oceans and sediments (Hayes et al., 1999). There is little isotopic discrimination effect during respiration but, during photosynthesis 12C is taken up preferentially
re
instead of 13C, leaving the superficial DIC pool enriched in 13C. As OM sinks into intermediate
lP
and deep waters, “sinking” organic carbon is remineralised and the deeper ocean DIC pool
na
becomes enriched in 12C. At geological time scales, removal of isotopically “light” OM into the sedimentary organic matter (SOM) reservoir generates a relative enrichment in 13C of dissolved
ur
CO2 in seawater and continuation of this process leads to an overall increase in 13C of marine
Jo
carbonates and OM (Jenkyns and Clayton, 1986; Wefer and Berger, 1991; Marshall, 1992; Patterson and Walter, 1994; Mitchell et al., 1996; Weissert et al., 1998; Holmden et al., 1998; Kump and Arthur, 1999; Hesselbo et al., 2000; Berner, 2003; Immenhauser et al., 2003; Swart and Eberli, 2005; Panchuk et al., 2006; Newton and Bottrell, 2007; Emerson and Hedges, 2008; Silva et al., 2011; Silva and Duarte, 2015).
1.2. Investigating oceanic carbon in the geological record
Journal Pre-proof Paired δ13Ccarb-TOC (δ13C (‰) = [(13C/12Csample)/(13C/12Cstandard)–1]*1000) analysis is one of the main tools to investigate perturbations of the carbon cycle at geological scales, where changes in δ13Ccarb and δ13CTOC are taken as proxies to changes in the 13C/12C ratio of OIC (dominated by DIC) and OOC (dominated by DPDOC, see above) respectively (Freeman and Hayes, 1992; Hayes et al., 1999; Berner, 2003) (Fig 1a). A conventional way to investigate the carbon cycle at geological time scales assumes that the carbon cycle is in (or near) steady state.
of
According to Krissansen-Totton et al. (2015), this assumption specifically means that i) the
ro
carbon input into surficial reservoirs (atmosphere, oceans, and biosphere) is balanced by the
-p
burial of carbonates plus organic carbon, and that ii) 13C/12C of inputs equal the average 13C/12C
re
of outputs (Schidlowski et al., 1979; Berner et al., 1983; Garrels and Lerman, 1984; Schidlowski, 1988; Beerling and Berner, 2002; Wallmann and Aloisi, 2012; Krissansen-Totton et al., 2015).
lP
This approach usually groups the atmosphere, oceans, life and soils, thus avoiding having to
na
explicitly include the short-term carbon cycle (Berner, 2004). This is a reasonable approximation for time intervals in the order of millions of years, but it may not hold when analysing high-
ur
resolution (hundreds to hundreds of thousands of years) sedimentary records (Sundquist, 1991;
Jo
Berner, 2004). At short-time scales, redistribution of carbon among the superficial reservoirs, likely driven by earth-orbital variations and internal ocean-atmosphere-biosphere-cryosphere feedbacks, may lead to “non-steady state” effects in the carbon cycle (Sundquist, 1985, 1991). For example, the observation of a temporal trend in the last 50 years in the Redfield ratio (carbon:nitrogen:phosphorus) of particle flux to the deep ocean suggests that the biological portion of the marine carbon cycle is not in steady state (Pahlow and Riebesell, 2000). In the geological record, Rothman et al. (2003) observed decoupling of the δ13Ccarb-TOC record during the Neoproterozoic. These authors suggested that in discrete periods of the
Journal Pre-proof Neoproterozoic, the carbon cycle was in non-steady state due to the accumulation of a large OOC pool (assumed to be at least 10x the size of the inorganic reservoir) and that the longer residence time of carbon in this reservoir buffered its isotopic response (i.e. slow equilibrium) to changes in δ13C of DIC. Rothman et al. (2003) also discussed that DOC may have dominated the OOC pool and that the large OOC reservoir lead to fluctuations in the isotopic composition of the smaller DIC reservoir either passively, by changing the isotopic composition of the fluxes
of
between DIC and OOC (ε0 not constant) or actively, due to enhanced organic remineralisation.
ro
More recently, it was proposed that the accumulation and remineralisation of oceanic
-p
DOC may have modulated the carbon cycle during the Paleocene and Eocene (Sexton et al.,
re
2011). Several intervals of rapid and pronounced global warming are recognised in this interval, including the Paleocene-Eocene Thermal Maximum (PETM). It was suggested that the less
lP
severe hyperthermals after the PETM resulted from astronomically paced large-scale releases of
na
DOC previously accumulated in anoxic areas due to ocean ventilation (Sexton et al., 2011). It was posited that the small isotopic fluctuations of about 1 ‰ associated with possible
ur
hyperthermals after the PETM resulted from the periodic accumulation and subsequent oxidation
Jo
of about 1600 Pg C from the DOC reservoir, approximately 2x the size of the DOC reservoir in today’s oceans (Hansell, 2013). Despite the uncertainty regarding, for example, the unknown sensitivity of refractory DOC degradation to ocean oxygenation, mechanisms for DOC accumulation, and oxidant availability (Bristow and Kennedy, 2008; Johnston et al., 2012; Ridgwell and Arndt, 2015), the studies of Rothman et al. (2003) and Sexton et al. (2011) call attention to the possibility of a nonsteady state carbon cycle and the possibility of OOC and DPOOC to act as a driver for planetaryscale changes at geological time scales.
Journal Pre-proof
1.3. Oceanic carbon as a driver for planetary-scale changes in the Mesozoic? The Bathonian–Callovian (168–163 Ma; Ogg et al., 2016), archives major and recurring sedimentary, biological, and climatic planetary-scale changes, including the hypothesised formation of polar ice, demise of carbonate platforms, eutrophication, high environmental stress, and several perturbations of the carbon cycle (Kenig et al., 1994; Hubbard and Boulter, 1997;
of
Price, 1999; Jenkyns et al., 2002; Dromart et al., 2003b, 2003a; Guex, 2006; O’Dogherty et al.,
ro
2006; Nunn et al., 2009; Andrieu et al., 2016; Koevoets et al., 2016). However, the drivers
-p
behind these planetary-scale changes and its consequences are still poorly understood (Dromart
re
et al., 2003b; Jenkyns et al., 2012; Arabas, 2016; Ikeda et al., 2016; Alberti et al., 2017; Song et al., 2017).
lP
We report here a paired δ13Ccarb-TOC record from the Upper Bathonian–Lower Callovian
na
hemipelagic succession cropping out at Praia da Pedra do Costado, Lusitanian Basin (Portugal). The δ13Ccarb-TOC record for this section is characterised by a short-lived decoupling event in the
ur
uppermost Bathonian–lowermost Callovian. The decoupling between δ13Ccarb-TOC suggests that
Jo
the short-term carbon cycle may have played a significant role in this event. In this paper, we examine potential causes for the δ13Ccarb-TOC decoupling event and their significance to the understanding of the oceanographic and environmental drivers that modulated the carbon cycle and marine dynamics in the Bathonian–Callovian.
2. Geological Background The Lusitanian Basin is a narrow and small North–South elongated basin, located in the western Iberian Margin (Wilson et al., 1989; Pinheiro et al., 1996; Rasmussen et al., 1998; Alves
Journal Pre-proof et al., 2002) (Fig. 2). The Lower to Middle Jurassic interval belongs to the first 1st-order sedimentary cycle of the Lusitanian Basin (Wilson et al., 1989; Alves et al., 2002), registered mostly by carbonate and shale deposits (Soares et al., 1993; Azerêdo et al., 2003). The temporal and lateral facies variation of theses successions highlight the basin evolution from a homoclinal ramp setting during the Lower Jurassic (e.g. Duarte et al., 2007; Silva et al., 2015) to a distally steepened ramp during the Middle Jurassic (Azerêdo, 1998; Azerêdo et al., 2002, 2014, 2020)
of
(Fig. 2). As the result of the regressive phase (started in the Lower Toarcian) of the 1st-order
ro
sedimentary cycle, two main sedimentary domains are differentiated in the Lusitanian Basin
-p
during the Middle Jurassic: to the east, the sedimentary series are typified by inner carbonate
re
ramp facies; to the western sectors, carbonate hemipelagic deposits crop out at Baleal and Cabo Mondego (Mangold, 1990; Azerêdo, 1998; Azerêdo et al., 2002, 2003, 2014; Canales and
lP
Henriques, 2008; Fernández-López et al., 2009) (Figs. 2 and 3).
na
The Cabo Mondego section, corresponding to the coastal areas of Cabo Mondego (Figueira da Foz, Portugal) contains the informal Cabo Mondego formation. This section is a
ur
worldwide stratigraphic reference that includes the Bajocian Global Boundary Stratotype Section
Jo
and Point (Pavia and Enay, 1997) and the Bathonian Auxiliary Section and Point (FernándezLópez et al., 2009). In the Cabo Mondego region, the Cabo Mondego formation encompasses the entire Middle Jurassic (e.g., Cariou et al., 1988; Mangold, 1990; Azerêdo et al., 2003; Fernández-López et al., 2009; Henriques et al., 2016) (Fig. 3). The Bathonian–Callovian boundary interval is particularly expressive at this location and is the only point in the basin where this boundary has some biostratigraphic control based on ammonites (Cariou et al., 1988; Mangold, 1990). Another point of interest is the occurrence of centimetric organic-rich shales dated from the Upper Bathonian, interpreted to represent the onset of a deepening phase and
Journal Pre-proof sedimentary starvation (Soares et al., 1993; Azerêdo et al., 2003) (Fig. 3). The Bathonian– Callovian deposits are interpreted as a 2nd-order transgressive-regressive facies cycle (Azerêdo et al., 2014). The studied portion of the Cabo Mondego section is located at Praia (beach) da Lage do Costado (29T; 507777.89 m E; 4448435.05 m N), Cabo Mondego, Figueira da Foz, Portugal (Fig. 2–4). The sampled and studied interval is dated from the Upper Bathonian–Lower
of
Callovian (Ruget-Perrot, 1961; Cariou et al., 1988; Mangold, 1990; Soares et al., 1993) and
ro
consist of about 40 metres of centimetric–decimetric shale(marl)–limestone alternations with a
-p
few centimetric apparently organic-rich beds exposed in a restricted wave-cut platform.
re
Ammonite macrofauna allows for a good biostratigraphic control of the Bathonian series. It was suggested (Cariou et al., 1988; Mangold, 1990) that the Bathonian–Callovian limit is located 8–
lP
13 metres above beds 152/153 (bed FN9 in Mangold, 1990). In our measured section, this
na
interval crudely corresponds to beds 180–210 (Fig. 4). The limit between the Middle–Upper Bathonian is poorly defined. Previous work
ur
(Mangold, 1990) have defined the Julii, Histricoides, and Angulicostatum subzones of the
Jo
Retrocostatum Zone in the Cabo Mondego section. According to (Mangold, 1990), the Retrocostatum Zone corresponds to the entire Upper Bathonian. The Julii Subzone corresponds to the lower and middle portion of the Orbis Zone, the Histricoides subzone to the upper Orbis Zone, and the Angulicostatum subzone to the uppermost Orbis Zone and Discus zone as defined by Torrens in (Cope et al., 1980) for the NW Europe. The remaining of the Lower Callovian section was inaccessible during the sampling campaign (see also Mangold, 1990), covered by sand, boulders, and cliff debris. The cliff section presented obvious signs of weathering (e.g. discolouration, karstification) and was avoided from sampling.
Journal Pre-proof
3. Materials and methods Seventy-four (74) samples were collected from the Upper Bathonian–Lower Callovian interval of the Cabo Mondego Formation in the Praia da Lage do Costado (Figs 2–4 and Table 1). From each sample, about 1-2 g of material was inspected for final cleaning, crushed, and then sent to Iso-Analytical Laboratories (United Kingdom) for determination of total organic carbon
of
(TOC), δ13C and δ18O in bulk carbonates (carbonates), and carbonate-free (13CTOC) isolates
ro
(Table 1).
-p
TOC was calculated first by obtaining % Cwhole-rock from the weight loss of the
re
decarbonated sample. The carbon content (total ion beam data) of the decarbonated sample (% Cdecarbonated) was measured on an elemental analyser (Europa Scientific Ltd) using IA-R001 wheat
lP
flour (40.2 %C w/w, Iso-Analytical Ltd, Crewe, UK) as the working reference material. The %
na
Cdecarbonated data were then corrected for the weight that was lost during the decarbonation procedure to calculate TOC (%). This method was validated against commercial TOC analysers
ur
(Iso-Analytical Ltd un-published data).
Jo
Stable carbon isotopes were determined in carbonate (13Ccarb) and carbonate-free (13CTOC) fractions. Stable oxygen isotopes were determined in carbonate (18Ocarb). All isotopic ratios are expressed relatively to VPDB (Coplen, 1994). The carbonate fractions were analysed by Continuous Flow–Isotope Ratio Mass Spectrometry (CF-IRMS). The samples were weighed into clean ExetainerTM tubes and then flushed with 99.995 % helium. After flushing, phosphoric acid was added to the samples, they were heated at 90ºC for 3 hours and then left to react in the acid overnight to allow complete conversion of carbonate to CO2. Reference and control materials were prepared the same way. CO2 was sampled into a continuously flowing He stream
Journal Pre-proof using a double holed needle and resolved on a packed column gas chromatograph and the resultant chromatographic peak carried forward into the ion source of a Europa Scientific 20-20 IRMS where it is ionized and accelerated. Gas species of different mass are separated in a magnetic field then simultaneously measured using a Faraday cup collector array (CO2 at m/z 44, 45, and 46). The reference materials used during analysis were IA-R022 (Iso-Analytical working standard calcium carbonate, 13C = -28.63 ‰ and 18O = -22.69 ‰), NBS-18 (carbonatite, 13C
of
= -5.01 ‰ and 18O = -23.2 ‰) and IA-R066 (chalk, 13C = +2.33 ‰ and 18O = -1.52 ‰). IA-
ro
R022 was calibrated against and is traceable to, NBS-18 and NBS-19 (limestone, δ13C = +1.95
-p
‰). IA-R066 was calibrated against and is traceable to NBS-18 and IAEA-CO-1 (Carrara
re
marble, δ13C = +2.5 ‰). NBS-18, NBS-19 and IAEA-CO-1 are inter-laboratory comparison standard materials distributed by the International Atomic Energy Agency (IAEA).
lP
For carbonate-free (13CTOC) isolates and Elemental Analysis- Isotope-Ratio Mass
na
Spectrometry (EA-IRMS) analysis, weighed sub-samples were taken from the sample vials, placed in universal tubes, acidified with 2M hydrochloric acid, mixed, oven heated at 60 °C for 2
ur
hours and left for 24 hours to allow all carbonate to be liberated as CO2. The sample fractions
Jo
were then isolated by centrifugation, and the acid was then decanted. The samples were then washed twice using distilled water and underwent centrifugation. After acid washing, the fractions were oven dried at 60 °C. Tin capsules containing a sample or reference material are loaded into an auto-sampler on a Europa Scientific elemental analyser, furnace held at 1000 °C and combusted in an oxygen-rich environment. The exhaust gas is reacted in a series of catalysts (e.g. chromate, silver wool) to oxidise hydrocarbons and remove sulphur and halides, purifying the CO2 stream. After purification, the remaining gas stream enters a Europa Scientific 20-20 IRMS ion source where it is ionised and accelerated. Gas species of different mass are separated
Journal Pre-proof in a magnetic field then simultaneously measured using a Faraday cup collector array (CO2 at m/z 44, 45, and 46). The reference material used during δ13CTOC analysis was IA-R001 (wheat flour, 13CV-PDB = of -26.43 ‰). For quality control purposes check samples of IA-R001, IAR005 (beet sugar, 13CV-PDB = -26.03 ‰) and IA-R006 (cane sugar, 13CV-PDB = -11.64 ‰) were analysed during batch analysis of the samples. IA-R001, IA-R005 and IA-R006 are calibrated against and traceable to IAEA-CH-6 (sucrose, 13CV-PDB = -10.43 ‰). IAEA-CH-6 is an inter-
of
laboratory comparison standard distributed by the International Atomic Energy Agency (IAEA).
ro
To ensure the precision of laboratory analyses, duplicated samples were introduced every
-p
five samples. Based on unpublished analyses and the data presented here, analytical precision is
lP
re
better than ±0.1‰.
4. Results
na
TOC in the studied section varies between 0.2–4.5 % (Fig 4 and Table 1). Most of the determined values are below 1 %. The highest TOC values are observed in black shales dated
ur
from the Histricoides Subzone, Retrocostatum Zone (Mangold, 1990).
Jo
δ13Ccarb varies from -0.45 to 3.16 ‰ (Fig 4 and Table 1). Overall, δ13Ccarb presents a slight negative trend for the ~10 m of the studied section dated from the Middle?-Upper Bathonian (Julii Subzone; Retrocostatum Zone (Mangold, 1990), followed by a positive trend, starting at the Julii–Histricoides subzone boundary through to the Lower Callovian (Mangold, 1990). Values of δ18Ocarb vary between -3.24 and -1.95 ‰ (Fig 4 and Table 1). Two δ18Ocarb maxima are observed; the first at the transition between the Julii and Histricoides subzones (Retrocostatum Zone) and the second at the Bathonian (Angulicostatum Subzone?)–Callovian
Journal Pre-proof boundary (Mangold, 1990). No significant correlation is observed between δ13Ccarb and δ18Ocarb (r2=0.16) (Fig 5). δ13CTOC varies from -27.39 to -23.18 ‰ (Fig 4 and Table 1). For the Bathonian, most of the δ13CTOC analyses range between around -27.3 to -26.5 ‰, decoupled from δ13Ccarb in the interval between the upper Bathonian (Retrocostatum Zone, Histricoides Subzone)–Lower Callovian (Mangold, 1990). The Lower Callovian succession records a sharp positive carbon
of
isotopic excursion (CIE) of about 1.3 ‰. Positive δ13CTOC values are also observed around the
ro
black shale intervals (Retrocostatum Zone, Histricoides Subzone (Mangold, 1990)), around 30 m
-p
from the base of the studied section. These are the most positive values of the studied section, ranging between -24 and -23 ‰. No significant correlation is observed between TOC and
re
δ13CTOC (r2=0.4252 considering the entire dataset, r2=0.172 without the high TOC (> 4 %)
na
lP
sample.
5. Diagenesis and preservation of a primary signal
ur
5.1. Are the δ13C and δ18O records of bulk carbonate a primary or diagenetic signal?
Jo
The effects of early marine diagenesis on carbonate δ13C and δ18O are not yet fully understood (Swart, 2015). However, and as a rule of thumb, inorganic aragonite and high-Mg calcite cements tend to have more positive δ13C than biologically derived skeletal aragonite (Braithwaite and Camoin, 2011). Early cementation influenced by-products linked with microbial OM decay, resulting in the production of CO2 depleted in 13C, may result in lower δ13C values (Marshall, 1992; Swart and Melim, 2000; Dickson et al., 2008; Coimbra et al., 2009; Silva et al., 2011; Duarte et al., 2014).
Journal Pre-proof The supply of oxidants controls the extent of the degradation of OM in sediments and it is apparent that under normal marine conditions oxidation of OM does not impact δ13C significantly in carbonate-rich rocks (Swart, 2015). On the other hand, the occurrence of diagenetic components influenced by OM oxidation might be significant in low or nondeposition intervals and where carbonates are a minor component (Swart, 2015). δ18O values of this kind of cements are also likely to vary depending on the nature of the original grains and the
of
temperature of precipitation. Precipitation of cement in cooler environments would result in
ro
more positive δ18O (Swart and Melim, 2000). Upon deeper burial, the temperature effect is
-p
further recorded on the carbonate δ18O, leading to lighter values due to the higher temperatures.
re
δ13C is more resistant to change during late-stage diagenesis as it has a very low thermal dependency and requires a high degree of water-rock interaction (Marshall, 1992; Jacobsen and
lP
Kaufman, 1999). However, the presence of high pCO2 and low δ13C fluids, a situation not
na
uncommon during late-stage diagenesis, can alter carbonate δ13C towards lighter values (Derry, 2010).
ur
The relatively carbonate-rich section and the low TOC contents of the studied samples
Jo
(generally <0.5 %) suggests that the δ13Ccarb trend reflects the primary oceanographic signal, even if the absolute values are slightly modified by diagenesis (Figs 1, 4, and 5). The lack of correlation between δ13Ccarb and δ18Ocarb supports the interpretation of the absence of a strong diagenetic imprint in the observed δ13Ccarb trend (Fig. 5). The correlation of our δ13Ccarb record with several contemporaneous datasets indicates that the observed temporal trend is, at least, a regional feature (Fig. 6).
5.2. Organic matter source and significance of the δ13CTOC profile from the Lusitanian Basin
Journal Pre-proof The obtained δ13CTOC record represents a mixed δ13C signal derived from primary production by aquatic organisms, microbial biomass (including dark carbon), and terrestrial material, all superimposed by diagenesis, thus resulting in potentially highly variable SOC δ13C signatures (Tyson, 1995; Emerson and Hedges, 2008; Herndl and Reinthaler, 2013; Hoefs, 2015; Suan et al., 2015) (Figs 1 and 4). Carbon isotopic composition of OM mainly depends on i) 13C content of the carbon source, including mixing; ii) isotope effects associated with the
of
assimilation of carbon; iii) isotope effects associated with metabolism and biosynthesis; and iv)
ro
cellular carbon budgets (Hayes et al., 1999; Hayes, 2001; Hoefs, 2015). Additional factors such
-p
as temperature, availability of CO2(aq), and nutrient availability were shown to exert some control on the δ13C of phytoplankton (Emerson and Hedges, 2008). Transformation of OM during its
re
transit in the water column seems not to impact significantly δ13C of organic carbon in
lP
sediments. However, chemotrophic addition to OM deposited in modern oxygen minimum zones
na
is shown to result in variably 13C depleted SOM (Rush et al., 2019) and the contribution of
(Galimov, 2006).
ur
reworked material makes SOM isotopically lighter (within 1–2‰) than its biological source
Jo
After deposition and during early diagenesis, two main processes change the isotopic composition of SOC: i) selected preservation of lipids (13C depleted) in relation to the more easily degradable carbohydrates and proteins, and ii) isotope fractionation due to the metabolism of microorganisms (Tyson, 1995; Hayes, 2001; Galimov, 2006; Hoefs, 2015). Preservation of lipid-rich 13C depleted marine SOM under oxygen-depleted conditions may produce negative δ13CTOC anomalies that do not necessarily reflect changes in the δ13C of the DIC pool (Suan et al., 2015). On the other hand, high-temperature thermal maturation processes and isotope exchange reactions between kerogen and carbonates may lead to a 13C enrichment of kerogen
Journal Pre-proof (Chung and Sackett, 1979; Lewan, 1983; Peters, 1986; Clayton, 1991; Hoefs, 2015), resulting in a shift of δ13CTOC to heavier values. From the above, it is irrefutable that a complex array of biological and diagenetic processes control δ13CTOC from any study. Despite its limitations, and keeping in mind that virtually all δ13CTOC records likely deviate from original δ13C, the overall shape and rates of change of δ13CTOC profiles are still of great utility to trace changes in paleoenvironmental and
of
paleoceanographic conditions contemporaneous with deposition (Oehlert and Swart, 2014; Jarvis
ro
et al., 2015; Suan et al., 2015).
-p
In well-preserved samples, the variation of OM composition is one of the main causes of δ13CTOC variation. We address this issue mainly by comparing the obtained δ13CTOC from
re
Portugal with published δ13C data from organic marine and terrestrial material. Upper Bajocian–
lP
Bathonian fossil wood from Yorkshire (UK) presents an average δ13C around -23 to -22 ‰ and
na
ranges between -24.2 to -21.6 ‰ (Hesselbo et al., 2003). In the mid-Callovian Oxford Clay Formation (UK), eleven (11) fossil wood samples have an average δ13C of -23.8 ‰ (Kenig et al.,
ur
1994). Kenig et al. (1994) estimated an average δ13C of -23.5 ‰ for terrestrial OM (δ13Cterrestrial)
Jo
and between -29.1 to -26.6 ‰ for primary marine OM in the Oxford Clay Formation. Based on the studies of Kenig et al. (1994) and Hesselbo et al. (2003), we assume a δ13Cterrestrial between 23.5 and -22.5 ‰ for the Bathonian–Callovian (nonetheless, further studies are needed to evaluate short-term δ13Cterrestrial variation for this time interval). The obtained δ13CTOC from the Upper Bathonian–Lower Callovian in the Lusitanian Basin is invariant for the most part, ranging between about -27.0 to -26.5 ‰ (Figs 4 and 6). The difference of 3–4 ‰ between Bathonian–Callovian δ13Cterrestrial and δ13CTOC from Portugal suggests a limited terrestrial contribution in the δ13CTOC record from the Lusitanian Basin in the
Journal Pre-proof decoupling interval. This observation is supported by another Mesozoic example suggesting that typical marine SOM usually yields δ13CTOC values that are 3–4‰ lighter when compared to coeval fossil wood or SOM of terrestrial origin (Tyson, 1995; Suan et al., 2015). δ13Cterrestrial is commonly interpreted to reflect the carbon isotopic composition of atmospheric CO2 and, consequently, δ13CDIC via air-water gas exchange (Gröcke et al., 1999; Gröcke, 2002; Hesselbo et al., 2007). Therefore, if terrestrial carbon was a major contributor to the studied δ13CTOC record
of
of the Lusitanian Basin, especially during the decoupling interval, the paired δ13Ccarb-TOC record
ro
should present some degree of co-variation. Finally, the invariant portion of the δ13CTOC curve in
-p
the decoupling interval (Figs 4 and 6) suggests a consistent and invariable isotopic composition
re
of SOC and likely similar SOM assemblages (Tyson, 1995). Summarizing, the δ13CTOC record from the Lusitanian Basin in the decoupling interval is interpreted to dominantly reflect a marine
lP
signal.
na
The interval between beds 140–166 (reaching up to about -24 ‰ in bed 164) and the last 5 m of the studied section (ranging between -26 to -25.5 ‰) present more positive δ13CTOC
ur
values. The first interval is dated from the Histricoides Subzone, Retrocostatum Zone, whereas
Jo
the second is dated from the Lower Callovian (Mangold, 1990) (Fig. 4). As discussed above, and assuming a terrestrial OM enrichment in 13C of about 3–4‰ in relation to marine OM, we postulate that the high-TOC laminated shales may have a significant terrestrial contribution, thus explaining the distinction between high-TOC laminated shales with heavier δ13CTOC vs low-TOC marls with lighter, and largely, invariant δ13CTOC (Fig 4). The δ13CTOC positive excursion of about 1‰ at the base of the Callovian is interpreted to reflect a change in the 13C/12C ratio or composition of SOM. The broad significance of the Late Bathonian–Early Callovian decoupling event and the Early Callovian δ13CTOC positive CIE is discussed in the following sections.
Journal Pre-proof
5.3. The Bathonian–Callovian regional δ13C record The regional δ13Ccarb profile for the Bathonian–Callovian includes the Early–Late Bathonian δ13Ccarb negative CIE, Late Bathonian–Early Callovian δ13Ccarb positive CIE, and the Early Callovian δ13Ccarb negative CIE (Kenig et al., 1994; Bartolini et al., 1995, 1996; Jenkyns, 1996; Gruszczyński, 1998; Bartolini et al., 1999; Jenkyns et al., 2002; O’Dogherty et al., 2006;
of
Nunn et al., 2009; Jach et al., 2014; Arabas, 2016; Koevoets et al., 2016; Alberti et al., 2017)
ro
(Fig. 6). In the well-dated La Cornicabra section (Spain), the Early–Late Bathonian δ13Ccarb
-p
negative CIE has an amplitude of about 0.5 ‰, the Late Bathonian–Early Callovian δ13Ccarb positive CIE has an amplitude of ~ 0.8 ‰, and the Early Callovian δ13Ccarb negative CIE has an
re
amplitude of about 0.9 ‰ (O’Dogherty et al., 2006) (Fig 6).
lP
Bathonian–Callovian δ13CTOC datasets are scarce and usually of very low temporal
na
resolution (Kenig et al., 1994; Nunn et al., 2009; Jach et al., 2014; Koevoets et al., 2016). Nunn et al. (2009) noted a decoupling event between δ13Ccarb and δ13CTOC during the Early Callovian in
ur
Scotland (within a lithological unit containing Kepplerites galilaei, Callomon in (Morton and
Jo
Hudson, 1995)) and recoupling in the Middle Callovian positive CIE. Nunn et al. (2009) stated: “it may relate to either a reduced exchange of carbon between the oceans and the atmosphere (for reasons unknown) or to the presence of local noise in the δ13C record”. This dataset is likely younger than the studied interval of the Lusitanian Basin. In Portugal, the δ13Ccarb-TOC minimum recorded around the Julii/Histricoides subzones (Retrocostatum Zone) is interpreted as the δ13C minimum interval of the Early–Late Bathonian δ13Ccarb negative CIE (Figs 4 and 6). The δ13Ccarb-TOC decoupling event is dated from the Late Bathonian (Retrocostatum Zone, Histricoides Subzone)–Early Callovian? (Mangold, 1990). The
Journal Pre-proof termination of the decoupling event is materialised by the Early Callovian δ13CTOC positive CIE, associated with the later stages of the Late Bathonian–Early Callovian δ13Ccarb positive CIE. In Portugal, the Late Bathonian–Early Callovian δ13Ccarb positive CIE has an amplitude of about 1.5–2 ‰ (Fig. 6). Based on available ammonite data from Portugal (Mangold, 1990), the δ13CcarbTOC
decoupling event is presumed to have lasted ~ 0.3 Myr (Ogg et al., 2016).
of
5.4. Possible mechanisms leading to the Late Bathonian–Early Callovian δ13Ccarb-TOC decoupling
ro
event in the Lusitanian Basin
-p
Covariant δ13Ccarb-TOC is usually interpreted as evidence that both carbonates and SOM have retained their original δ13C composition (Oehlert and Swart, 2014). On the other hand,
re
decoupled δ13Ccarb-TOC records are much more infrequent and usually remain unexplained (Nunn
lP
et al., 2009; Yager et al., 2017) or are taken to reflect i) diagenetic alteration (Silva et al., 2013),
na
ii) detrital flux of exogenous (fossil) OM or migration of hydrocarbons (Johnston et al., 2012), iii) local changes in DIC or DOC (Patterson and Walter, 1994; Immenhauser et al., 2003; Suan et
ur
al., 2015), or iv) the Rothman ocean model (changes in size of the OOC pool and non-steady
Jo
state C cycle, see the Introduction section) Rothman et al., 2003) and variations thereof (Fike et al., 2006; Peltier et al., 2007; Swanson-Hysell et al., 2010). As discussed previously, the temporal evolution of the Upper Bathonian–Lower Callovian paired δ13Ccarb-TOC record from the Lusitanian Basin is here interpreted to reflect environmental, oceanographic, and diagenetic conditions and variation contemporaneous with deposition. The lack of correlation between δ13Ccarb and δ18Ocarb (Fig. 5) and the similarity of our δ13Ccarb record with several contemporaneous datasets suggest a weak diagenetic imprint in the δ13Ccarb record and a temporal, at least, regional variation pattern (Fig. 6). Extensive diagenetic
Journal Pre-proof alteration or terrestrial “contamination” are discarded as major mechanisms driving the δ13CTOC trend (see section 5.2). Explaining the observed invariant δ13CTOC record using the basinal OM vs exogenous OM model (Johnston et al., 2012) also seems unlikely. To our knowledge, there is no evidence of an exogenous source of organic carbon in the Lusitanian Basin, either from the erosion of organic-rich shales on land or hydrocarbons seepage. Provokingly, and as a working hypothesis, we suggest that the δ13Ccarb-TOC decoupling
of
event and the Bathonian–Early Callovian CIEs resulted from the addition and removal of carbon
ro
from the OOC reservoir, likely DPOOC (Figs 1, 6, and 7). A similar mechanism can be evoked
-p
to explain the likely younger decoupling event and the Middle Callovian δ13Ccarb positive CIE in
re
Nunn et al. (2009). As in Rothman et al. (2003), it is speculated that the mostly invariant δ13CTOC record during the decoupling event resulted from the accumulation of a large OOC reservoir,
lP
which increased the residence time of carbon in the oceans and buffered δ13CTOC (slow
na
equilibrium) to changes in δ13CDIC (see Ridgwell and Arndt (2015) regarding the limitations of this interpretation). The relative size, chemical and isotopic composition, and oceanographical
ur
distribution of the DPOOC and DIC pools and composition and magnitude of the fluxes between
Jo
the different superficial carbon reservoirs, including external inputs, are unknown (Figs 1 and 7) and limit further exploitation of proposed process in the Mesozoic. Investigation of contemporaneous sections from other geographical location, development of methods for a complete and detailed characterisation of SOM, and development of carbon cycle models that explicitly include OOC and DPOOC as a readily exchangeable superficial carbon reservoir are needed to constrain spatially and temporally the Bathonian–Callovian intervals. The accumulation of OOC implies a non-steady stated of the short-term carbon cycle in the Bathonian–Callovian.
Journal Pre-proof However, and regardless of the Bathonian–Callovian oceans ability to accumulate the vast amounts of OOC required to effectively buffer δ13CTOC (see Ridgwell and Arndt (2015) regarding practical limitations of this interpretation), the temporal variation in the size of a more “modern-sized” OOC reservoir may have had impacted δ13Ccarb and δ13CTOC, either via oxidation of DPOOC to CO2 or enhanced burial of organic matter from OOC (Rothman et al., 2003; Fike et al., 2006; Sexton et al., 2011; Ridgwell and Arndt, 2015). In the following paragraphs, it is
of
assumed that the accumulation of OOC and DPOOC occurred in a large area of the global ocean.
ro
A testable prediction of our working hypothesis concerns the fate of the DPOOC pool
-p
and its impact on the global carbon cycle and climate: if part of this reservoir is oxidized to CO2, increases in atmospheric pCO2 and temperature are likely to occur and a negative CIE and
re
recoupling of δ13Ccarb-TOC are expected; if DPOOC is exported to sediments, generalized organic
lP
matter deposition may occur and a positive CIE and recoupling of δ13Ccarb-TOC are expected (see
na
also (Rothman et al., 2003; Fike et al., 2006; Bekker et al., 2008; Sexton et al., 2011; Ridgwell and Arndt, 2015)) (Fig. 7). The Early–Late Bathonian δ13Ccarb negative CIE (about 0.5 ‰ in the
ur
La Cornicabra section, about 1.5–2 ‰ in Portugal) and the Early Callovian δ13Ccarb negative CIE
Jo
(about 0.9 ‰ in the La Cornicabra section) are here explained as resulting from the oxidation of DPOOC to CO2 (Fig. 6 and 7). The causes for the different amplitudes of the δ13Ccarb negative CIE between Portugal and Spain may be related to the poor resolution of the La Cornicabra section, diagenesis artefacts, spatial heterogeneity of the OIC and OOC reservoirs, different carbonate sources, or local factors affecting δ13Ccarb. Using the modern ocean carbon inventory as a reference and a very simplistic mass calculation exercise with an initial isotopic composition of δ13CDIC of 2 ‰ and an DPOOC source isotopic signature of -27 ‰ (eq 1.3b in Ridgwell and Arndt, 2015), negative CIEs with amplitudes of 0.5, 1, and 2 ‰ can be explained by the
Journal Pre-proof oxidation of about 700, 1500, and 3000 Pg of C from DPOOC respectively. These quantities represent roughly one to four times the modern-day DOC inventory (Hansell, 2013). As predicted, and despite the lack of detailed records, the two negative CIEs are apparently associated with relatively higher paleotemperatures (Dera et al., 2011). The Early Callovian δ13CTOC positive CIE and recoupling with δ13Ccarb (coeval with δ13Ccarb maximum of the latest Bathonian–Early Callovian δ13Ccarb positive CIE, O’Dogherty et
of
al., 2006) is interpreted to be the result of increased efficiency in OOC and DPOOC export into
ro
sediments, materialized by the generalized deposition of organic matter during the Early
-p
Callovian (Fig. 6 and 7). Lower Callovian marine and terrestrial organic-rich facies and black
re
shales are known from Spain (Bullatus Zone, Lower Callovian (Gräfe, 2005)), Northeast Scotland (Macrocephalus zone, Lower Callovian (Nagy et al., 2001)), central North America
lP
(Lower Callovian (Imlay, 1981)), Himalayas (Discus?–Gracilis zones; Upper Bathonian–Lower
na
Callovian (Yin, 2007)), east Greenland (Lower Callovian (Callomon, 1993)), Atlantic Canada (Lower Callovian (Mukhopadhyay and Wade, 1990)), west-Gulf of Mexico (Cantú–Chapa,
ur
1969) and Japan (Kusuhashi et al., 2002; Handa et al., 2014).
Jo
The wax and wane of the DPOOC component of the OOC reservoir are here viewed as the redistribution of carbon among Earth’s superficial reservoirs, driven by earth-orbital forcing mechanisms and internal ocean-atmosphere-biosphere-cryosphere feedbacks (Sundquist, 1985, 1991) and interaction with geosphere. Currently, there is a large uncertainty regarding sources and processes of accumulation and removal of recalcitrant DOC, which comprises the majority of the OOC reservoir in today’s oceans (Hansell, 2013; Follett et al., 2014; Legendre et al., 2015; Polimene et al., 2017; Wilson and Arndt, 2017). Examples of important sources of DOC include terrestrial contributions, viral lysis, phytoplankton release, “sloppy feeding” by metazoan
Journal Pre-proof grazers, egesta of protists and metazoan, particulate organic matter solubilization by bacterial and archeal ectohydrolases (Jiao et al., 2010; Hansell, 2013; Carlson and Hansell, 2015; Legendre et al., 2015). Oceanic accumulation of recalcitrant DOC is thought to occur by abiotic processes (via condensation reactions catalysed by light and metal complexation, polymerization, or adsorption to colloids) and biologically, mainly associated with the microbial carbon pump, limitation of heterotrophic bacterioplankton growth (via nutrient limitation, microbial
of
community, pressure, and temperature), and eukaryotic production; the main processes of
ro
removal include biotic consumption by heterotrophic prokaryotes (via bacterioplankton
-p
consumption or remineralization) and eukaryotes, photochemical oxidation of recalcitrant DOC
re
via ocean mixing, sorption of DOM onto sinking particles, condensation of gels, and hydrothermal circulation (see, for example, (Druffel et al., 1992; Jiao et al., 2010, 2014; Carlson
lP
et al., 2011; Arndt et al., 2013; Hansell, 2013; Follett et al., 2014; Legendre et al., 2015;
na
Ridgwell and Arndt, 2015; Carlson and Hansell, 2015; Dittmar, 2015; Cael et al., 2017; Wilson and Arndt, 2017) and references therein).
ur
Any of the aforementioned processes, or a combination thereof, might have modulated
Jo
the changes in size of the OOC reservoir and the DPOOC component in the Bathonian– Callovian. For example, it was suggested that ocean temperature changes might affect the efficiency of the biological carbon pump and the partitioning of organic carbon between POC and DOC, with warming shifting the partitioning between POC and DOC toward enhanced accumulation of dissolved compounds (Wohlers et al., 2009; Legendre et al., 2015). It seems reasonable to extrapolate that astronomically paced variation of several paleoclimatic parameters may have had resulted in changes of the relative efficiency of the biological carbon pump and microbial carbon pump and in the postulated changes in the sizes of the several superficial
Journal Pre-proof carbon reservoirs (Fig. 7). In addition to the effect of temperature in the biological carbon pump, astronomically paced variation in insolation might have also controlled the overall rates of photochemical oxidation of recalcitrant DOC and nutrient availability via climate and weathering. Even if we are still distant from a reliable model capable of describing the variations in size of the different oceanic and atmospheric carbon reservoirs in short-lived geological
of
intervals, the working hypothesis of accumulation and remineralization of OOC and DPOOC as
ro
a first-order control on the carbon cycle and climate during specific intervals of the Mesozoic
-p
provides a testable hypothesis to explain the astronomically paced short-lived ~1 ‰ negative
re
CIEs that infuse the Mesozoic, especially the small fluctuations within larger-scale CIEs associated with oceanic anoxic events (Schlanger and Jenkyns, 1976; Hesselbo et al., 2007).
lP
Small-scale isotopic perturbations that might be explained by the OOC working hypothesis can
na
be found, for example, in the Early Jurassic Liasidium Event (Hesselbo et al., 2019), Early Toarcian oceanic anoxic event (OAE) (Hesselbo et al., 2007; Boulila et al., 2014; Thibault et al.,
ur
2018), or Late Aptian–Early Albian OAE 1b cluster (Herrle et al., 2004; Föllmi et al., 2006). If
Jo
OOC, either alone or combined with additional inputs from the terrestrial biosphere, marine hydrates, and/or enhanced volcanic outgassing, can explain larger amplitude Mesozoic CIEs must also be qualitatively explored (Ridgwell and Arndt, 2015).
6. Conclusions and final remarks In this study, we present evidence of at least one δ13Ccarb-TOC decoupling event in the Bathonian–Callovian. Provokingly, we suggest that this event resulted from the accumulation of a large pool of oceanic organic carbon (OOC), likely dominated by dissolved and fine particulate
Journal Pre-proof oceanic organic carbon. Expansion of the OOC reservoir increased the residence time of carbon in the oceans and buffered δ13COOC (approximated by δ13CTOC) to changes in δ13C of oceanic inorganic carbon, dominated by dissolved inorganic carbon (approximated by δ13Ccarb). Variation in the size of the OOC pool and redistribution of carbon among the superficial carbon reservoirs provides an alternative explanation for the carbon isotopic excursions (CIEs) observed in the Bathonian–Callovian; oxidation of OOC to CO2 resulted in small-scale negative CIEs (~ 1‰),
of
whereas the change from OOC production and accumulation to carbon export to sediments lead
ro
to positive CIEs. The reduction in size of the OOC pool lead to the recoupling of the δ13Ccarb-TOC
-p
records. Other small-scale (~ 1‰) δ13Ccarb-TOC excursions in the Mesozoic may have resulted
re
from similar episodes of OOC accumulation and oxidation.
We hope that our bold claims stimulate further research in this often-overlooked interval
lP
and highlight the need to explicitly consider OOC (especially DOC) in current carbon cycle
na
modelling efforts of high-resolution geological records. The working conceptual model presented here for the Bathonian-Callovian is readily testable by the acquisition of other detailed
ur
paired δ13Ccarb-TOC records and comprehensive characterization of organic matter source from
Jo
other basins and detailed modelling of the carbon cycle with an explicit consideration of the short-term cycle and its major carbon reservoirs.
Acknowledgements This research was supported by the Source Rock and Geochemistry of the Central Atlantic Margins consortium (Dalhousie University, Basin and Reservoir Lab, PI – G. Wach). L. V. Duarte was supported by FCT, through the strategic project UID/MAR/04292/2019 granted to the Marine and Environmental Sciences Centre (MARE, Portugal). R. L. Silva was also partially
Journal Pre-proof supported by the Irish Centre for Research in Applied Geosciences, iCRAG (project: Temporal and spatial variability in Lower Jurassic hydrocarbon source rock quality in Irish off-shore marine basins, PI - Micha Ruhl). We also thank Katja Fennel, Stephen P. Hesselbo and Micha Ruhl for helpful scientific discussions on some of the topics covered in this manuscript; and the Editor-in-Chief Fabienne Marret-Davies and two anonymous referees for their insightful
Jo
ur
na
lP
re
-p
ro
of
comments that greatly benefited the manuscript.
Journal Pre-proof References Alberti, M., Fürsich, F.T., Abdelhady, A.A., Andersen, N., 2017. Middle to Late Jurassic equatorial seawater temperatures and latitudinal temperature gradients based on stable isotopes of brachiopods and oysters from Gebel Maghara, Egypt. Palaeogeogr. Palaeoclimatol. Palaeoecol. 468, 301–313. https://doi.org/10.1016/j.palaeo.2016.11.052 Alves, T.M., Gawthorpe, R.L., Hunt, D.W., Monteiro, J.H., 2002. Jurassic tectono-sedimentary
ro
754. https://doi.org/10.1016/S0264-8172(02)00036-3
of
evolution of the Northern Lusitanian Basin (offshore Portugal). Mar. Pet. Geol. 19, 727–
-p
Andrieu, S., Brigaud, B., Barbarand, J., Lasseur, E., Saucède, T., 2016. Disentangling the control
re
of tectonics, eustasy, trophic conditions and climate on shallow-marine carbonate production during the Aalenian–Oxfordian interval: From the western France platform to
lP
the western Tethyan domain. Sediment. Geol. 345, 54–84.
na
https://doi.org/10.1016/j.sedgeo.2016.09.005 Arabas, A., 2016. Middle-Upper Jurassic stable isotope records and seawater temperature
ur
variations: New palaeoclimate data from marine carbonate and belemnite rostra (Pieniny
Jo
Klippen Belt, Carpathians). Palaeogeogr. Palaeoclimatol. Palaeoecol. 446, 284–294. https://doi.org/10.1016/j.palaeo.2016.01.033 Arndt, S., Jørgensen, B.B., LaRowe, D.E., Middelburg, J.J., Pancost, R.D., Regnier, P., 2013. Quantifying the degradation of organic matter in marine sediments: A review and synthesis. Earth-Science Rev. 123, 53–86. https://doi.org/10.1016/j.earscirev.2013.02.008 Azerêdo, A.C., 1998. Geometry and facies dynamics of Middle Jurassic carbonate ramp sandbodies, West-Central Portugal. Geol. Soc. London, Spec. Publ. 149, 281–314. https://doi.org/10.1144/GSL.SP.1999.149.01.14
Journal Pre-proof Azerêdo, A.C., Duarte, L.V., Henriques, M.H., Manuppella, G., 2003. Da dinâmica continental no Triásico aos mares do Jurássico Inferior e Médio. Cadernos de Geologia de Portugal. Instituto Geológico e Mineiro. Azerêdo, A.C., Duarte, L.V., Silva, R.L., 2014. Sequential configuration in transgressiveregressive facies cycles (2nd order) of the Lusitanian Basin (Portugal) Lower and Middle Jurassic. Comun. Geol. 101, 383–386.
of
Azerêdo, A.C., Inês, N., Bizarro, P., 2020. Carbonate reservoir outcrop analogues with a glance
-p
https://doi.org/10.1016/j.marpetgeo.2019.08.034
ro
at pore-scale (Middle Jurassic, Lusitanian Basin, Portugal). Mar. Pet. Geol. 111, 815–851.
re
Azerêdo, A.C., Wright, V.P., Ramalho, M.M., Azeredo, A.C., Wright, V.P., Ramalho, M.M., 2002. The Middle-Late Jurassic forced regression and disconformity in central Portugal:
lP
Eustatic, tectonic and climatic effects on a carbonate ramp system. Sedimentology 49,
na
1339–1370. https://doi.org/10.1046/j.1365-3091.2002.00501.x Bartolini, A., Baumgartner, P.O., Guex, J., 1999. Middle and Late Jurassic radiolarian
ur
palaeoecology versus carbon-isotope stratigraphy. Palaeogeogr. Palaeoclimatol. Palaeoecol.
Jo
145, 43–60. https://doi.org/10.1016/S0031-0182(98)00097-2 Bartolini, A., Baumgartner, P.O., Hunziker, J., 1996. Middle and Late Jurassic carbon stableisotope stratigraphy and radiolarite sedimentation of the Umbria-Marche basin (central Italy). Eclogae Geol. Helv. 89, 811–844. https://doi.org/10.5169/seals-167925 Bartolini, A., Baumgartner, P.O., Mattioli, E., 1995. Middle and Late Jurassic radiolarian biostratigraphy of the Colle Bertone and Terminilletto sections (Umbria-Marche-Sabina Appennines, Central Italy): an integrated stratigraphy approach, in: Baumgartner, P.O. et al (Ed.), Middle Jurassic to Lower Cretaceous Radiolaria of Tethys: Occurrence, Systematics,
Journal Pre-proof Biochronology. InterRad Jurassic-Cretaceous Working Group,23, pp. 817–831. Beerling, D.J., Berner, R.A., 2002. Biogeochemical constraints on the Triassic-Jurassic boundary carbon cycle event. Global Biogeochem. Cycles 16, 10-1-10–13. https://doi.org/10.1029/2001GB001637 Bekker, A., Holmden, C., Beukes, N.J., Kenig, F., Eglinton, B., Patterson, W.P., 2008. Fractionation between inorganic and organic carbon during the Lomagundi (2.22–2.1 Ga)
of
carbon isotope excursion. https://doi.org/10.1016/j.epsl.2008.04.021
ro
Berner, R. a, Lasaga, a C., Garrels, R.M., 1983. The carbonate-silicate geochemical cycle and its
re
683. https://doi.org/10.2475/ajs.283.7.641
-p
effect on atmospheric carbon dioxide over the past 100 million years. Am. J. Sci. 283, 641–
Berner, R.A., 2006. GEOCARBSULF: A combined model for Phanerozoic atmospheric O2 and
lP
CO2. Geochim. Cosmochim. Acta 70, 5653–5664.
na
https://doi.org/10.1016/j.gca.2005.11.032
Berner, R.A., 2004. The Phanerozoic Carbon Cycle: CO2 and O2. Oxford University Press.
ur
Berner, R.A., 2003. The long-term carbon cycle, fossil fuels and atmospheric composition.
Jo
Nature. https://doi.org/10.1038/nature02131 Berner, R.A., 1999. A new look at the long-term carbon cycle. GSA Today 9, 2–6. https://doi.org/10.1002/qj.49710443903 Boulila, S., Galbrun, B., Huret, E., Hinnov, L.A., Rouget, I., Gardin, S., Bartolini, A., 2014. Astronomical calibration of the Toarcian Stage: Implications for sequence stratigraphy and duration of the early Toarcian OAE. Earth Planet. Sci. Lett. 386, 98–111. https://doi.org/10.1016/j.epsl.2013.10.047 Braithwaite, C.J.R., Camoin, G.F., 2011. Diagenesis and sea-level change: lessons from
Journal Pre-proof Moruroa, French Polynesia. Sedimentology 58, 259–284. https://doi.org/10.1111/j.13653091.2010.01182.x Bristow, T.F., Kennedy, M.J., 2008. Carbon isotope excursions and the oxidant budget of the Ediacaran atmosphere and ocean. Geology. https://doi.org/10.1130/G24968A.1 Cael, B.B., Bisson, K., Follows, M.J., 2017. How have recent temperature changes affected the efficiency of ocean biological carbon export? Limnol. Oceanogr. Lett.
of
https://doi.org/10.1002/lol2.10042
ro
Callomon, J.H., 1993. The ammonite succession in the Middle Jurassic of East Greenland. Bull.
-p
Geol. Soc. Denmark 40, 83–113.
re
Canales, M.L., Henriques, M.H., 2008. Foraminifera from the Aalenian and the Bajocian GSSP (Middle Jurassic) of Murtinheira section (Cabo Mondego, West Portugal): Biostratigraphy
lP
and paleoenvironmental implications. Mar. Micropaleontol. 67, 155–179.
na
https://doi.org/10.1016/j.marmicro.2008.01.003 Cantú–Chapa, A., 1969. Estratigrafía del Jurásico Medio–Superior del subsuelo de Poza Rica,
ur
Ver. (Área de Soledad–Miquetla). Rev. del Inst. Mex. del Petróleo 1, 3–9.
Jo
Cariou, E., Mangold, C., Thierry, J., Ruget, C., Mouterde, R., Rocha, R.B., 1988. Biochronologie du Callovien inferieur et moyen de la province submediterraneenne: apport de la coupe du Cap Mondego (Portugal), in: 2nd Int. Symposium Jurassic Stratigraphy. pp. 407–418. Carlson, C.A., Hansell, D.A., 2015. DOM Sources, Sinks, Reactivity, and Budgets, in: Hansell, D.A., Carlson, C.A. (Eds.), Biogeochemistry of Marine Dissolved Organic Matter. Elsevier, pp. 65–126. https://doi.org/10.1016/B978-0-12-405940-5.00003-0 Carlson, C.A., Hansell, D.A., Tamburini, C., 2011. DOC persistance and its fate after export within the ocean interior, in: Microbial Carbon Pump in the Ocean. pp. 57–59.
Journal Pre-proof https://doi.org/doi: 10.1126/science.opms.sb0001 Cavan, E.L., Trimmer, M., Shelley, F., Sanders, R., 2017. Remineralization of particulate organic carbon in an ocean oxygen minimum zone. Nat. Commun. 8, 14847. https://doi.org/10.1038/ncomms14847 Chung, H.M., Sackett, W.M., 1979. Use of stable carbon isotope compositions of pyrolytically derived methane as maturity indices for carbonaceous materials. Geochim. Cosmochim.
of
Acta 43, 1979–1988. https://doi.org/10.1016/0016-7037(79)90010-3
ro
Ciais, P., Sabine, C., Bala, G., Bopp, L., Brovkin, V., Canadell, J., Chhabra, A., DeFries, R.,
-p
Galloway, J., Heimann, M., Jones, C., Quéré, C. Le, Myneni, R.B., Piao, S., Thornton, P.,
re
2013. Carbon and Other Biogeochemical Cycles, in: Stocker, T.F., Qin, D., Plattner, G.-K., Tignor, M., Allen, S.K., Boschung, J., Nauels, A., Xia, Y., Bex, V., Midgley, P.M. (Eds.),
lP
Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the
na
Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 465–570.
ur
Clayton, C., 1991. Carbon isotope fractionation during natural gas generation from kerogen.
Jo
Mar. Pet. Geol. 8, 232–240. https://doi.org/10.1016/0264-8172(91)90010-X Coimbra, R., Immenhauser, A., Olóriz, F., 2009. Matrix micrite δ13C and δ18O reveals synsedimentary marine lithification in Upper Jurassic Ammonitico Rosso limestones (Betic Cordillera, SE Spain). Sediment. Geol. 219, 332–348. https://doi.org/10.1016/j.sedgeo.2009.06.002 Cope, J.C.W., Duff, K.L., Parsons, C.F., Torrens, H.S., Wimbledon, W.A., K, W.J., 1980. A correlation of Jurassic rocks in the British Isles: Part Two: Middle and Upper Jurassic, Geological Society of London, Special Report No. 15. The Geological Society, London.
Journal Pre-proof Coplen, T.B., 1994. Reporting of stable hydrogen, carbon, and oxygen isotopic abundances (Technical Report). Pure Appl. Chem. 66, 273–276. https://doi.org/10.1351/pac199466020273 Dera, G., Brigaud, B., Monna, F., Laffont, R., Pucéat, E., Deconinck, J.-F., Pellenard, P., Joachimski, M.M., Durlet, C., 2011. Climatic ups and downs in a disturbed Jurassic world. Geology 39, 215–218. https://doi.org/10.1130/G31579.1
of
Derry, L.A., 2010. A burial diagenesis origin for the Ediacaran Shuram–Wonoka carbon isotope
ro
anomaly. Earth Planet. Sci. Lett. 294, 152–162. https://doi.org/10.1016/j.epsl.2010.03.022
-p
Dickson, J.A.D., Wood, R.A., Bu Al Rougha, H., Shebl, H., 2008. Sulphate reduction associated
re
with hardgrounds: Lithification afterburn! Sediment. Geol. 205, 34–39. https://doi.org/10.1016/j.sedgeo.2008.01.005
lP
Dittmar, T., 2015. Reasons Behind the Long-Term Stability of Dissolved Organic Matter, in:
na
Hansell, D.A., Carlson, C.A. (Eds.), Biogeochemistry of Marine Dissolved Organic Matter. Elsevier, pp. 369–388. https://doi.org/10.1016/B978-0-12-405940-5.00007-8
ur
Dromart, G., Garcia, J.-P., Gaumet, F., Picard, S., Rousseau, M., Atrops, F., Lecuyer, C.,
Jo
Sheppard, S.M.F., 2003a. Perturbation of the carbon cycle at the Middle/Late Jurassic transition: Geological and geochemical evidence. Am. J. Sci. 303, 667–707. https://doi.org/10.2475/ajs.303.8.667 Dromart, G., Garcia, J.-P., Picard, S., Atrops, F., Lécuyer, C., Sheppard, S., 2003b. Ice age at the Middle–Late Jurassic transition? Earth Planet. Sci. Lett. 213, 205–220. https://doi.org/10.1016/S0012-821X(03)00287-5 Druffel, E.R.M., Williams, P.M., Bauer, J.E., Ertel, J.R., 1992. Cycling of dissolved and particulate organic matter in the open ocean. J. Geophys. Res. 97, 15639.
Journal Pre-proof https://doi.org/10.1029/92JC01511 Duarte, L.V., Comas-Rengifo, M.J., Silva, R.L., Paredes, R., Goy, A., 2014. Carbon isotope stratigraphy and ammonite biochronostratigraphy across the Sinemurian-Pliensbachian boundary in the western Iberian margin. Bull. Geosci. 89, 719–736. https://doi.org/10.3140/bull.geosci.1476 Duarte, L. V., Oliveira, L.C., Rodrigues, R., 2007. Carbon isotopes as a sequence stratigraphic
of
tool: Examples from the Lower and Middle Toarcian marly limestones of Portugal. Bol.
ro
Geol. y Min. 118, 3–18.
-p
Emerson, S., Hedges, J., 2008. Chemical oceanography and the marine carbon cycle. Cambridge
re
University Press.
Fernández-López, S.R., Pavia, G., Erba, E., Guiomar, M., Henriques, M.H., Lanza, R., Mangold,
lP
C., Morton, N., 2009. The Global Boundary Stratotype Section and Point (GSSP) for base
222–248.
na
of the Bathonian Stage (Middle Jurassic), Ravin du Bès Section , SE France. Episodes 32,
ur
Fike, D.A., Grotzinger, J.P., Pratt, L.M., Summons, R.E., 2006. Oxidation of the Ediacaran
Jo
Ocean. Nature 444, 744–747. https://doi.org/10.1038/nature05345 Follett, C.L., Repeta, D.J., Rothman, D.H., Xu, L., Santinelli, C., 2014. Hidden cycle of dissolved organic carbon in the deep ocean. Proc. Natl. Acad. Sci. U. S. A. 111, 16706– 16711. https://doi.org/10.1073/pnas.1407445111 Föllmi, K.B., Godet, A., Bodin, S., Linder, P., 2006. Interactions between environmental change and shallow water carbonate buildup along the northern Tethyan margin and their impact on the Early Cretaceous carbon isotope record. Paleoceanography 21. https://doi.org/10.1029/2006PA001313
Journal Pre-proof Freeman, K.H., Hayes, J.M., 1992. Fractionation of carbon isotopes by phytoplankton and estimates of ancient CO 2 levels. Global Biogeochem. Cycles 6, 185–198. https://doi.org/10.1029/92GB00190 Galimov, E.M., 2006. Isotope organic geochemistry. Org. Geochem. 37, 1200–1262. https://doi.org/10.1016/j.orggeochem.2006.04.009 Garrels, R.M., Lerman, A., 1984. Coupling of the sedimentary sulfur and carbon cycles; an
of
improved model. Am. J. Sci. 284, 989–1007. https://doi.org/10.2475/ajs.284.9.989
ro
Gräfe, K.-U., 2005. Benthic foraminifers and palaeoenvironment in the Lower and Middle
-p
Jurassic of the Western Basque-Cantabrian Basin (Northern Spain). J. Iber. Geol. 31, 217–
re
233.
Gröcke, D.R., 2002. The carbon isotope composition of ancient CO2 based on higher-plant
lP
organic matter. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 360, 633–658.
na
https://doi.org/10.1098/rsta.2001.0965
Gröcke, D.R., Hesselbo, S.P., Jenkyns, H.C., 1999. Carbon-isotope composition of Lower
ur
Cretaceous fossil wood: Ocean-atmosphere chemistry and relation to sea-level change.
Jo
Geology 27, 155. https://doi.org/10.1130/0091-7613(1999)027<0155:CICOLC>2.3.CO;2 Gruszczyński, M., 1998. Chemistry of Jurassic seas and its bearing on the existing organic life. Acta Geol. Pol. 161, 1–29. Guex, J., 2006. Reinitialization of evolutionary clocks during sublethal environmental stress in some invertebrates. Earth Planet. Sci. Lett. 242, 240–253. https://doi.org/10.1016/j.epsl.2005.12.007 Handa, N., Nakada, K., Anso, J., Matsuoka, A., 2014. Bathonian/Callovian (Middle Jurassic) ammonite biostratigraphy of the Kaizara Formation of the Tetori Group in central Japan.
Journal Pre-proof Newsletters Stratigr. 47, 283–297. https://doi.org/10.1127/0078-0421/2014/0048 Hansell, D.A., 2013. Recalcitrant Dissolved Organic Carbon Fractions. Ann. Rev. Mar. Sci. 5, 421–445. https://doi.org/10.1146/annurev-marine-120710-100757 Hansell, D.A., Carlson, C.A., 2013. Localized refractory dissolved organic carbon sinks in the deep ocean. Global Biogeochem. Cycles 27, 705–710. https://doi.org/10.1002/gbc.20067
Rev. Mineral. Geochemistry 43, 225–277.
ro
https://doi.org/https://doi.org/10.2138/gsrmg.43.1.225
of
Hayes, J.M., 2001. Fractionation of carbon and hydrogen isotopes in biosynthetic processes.
-p
Hayes, J.M., Strauss, H., Kaufman, A.J., 1999. The abundance of 13C in marine organic matter
re
and isotopic fractionation in the global biogeochemical cycle of carbon during the past 800 Ma. Chem. Geol. 161, 103–125. https://doi.org/10.1016/S0009-2541(99)00083-2
lP
Henriques, M.H., Canales, M.L., Silva, S.C., Figueiredo, V., 2016. Integrated Biostratigraphy
na
(Ammonoidea, Foraminiferida) of the Aalenian of the Lusitanian Basin (Portugal): A synthesis. Episodes 39, 482. https://doi.org/10.18814/epiiugs/2016/v39i3/99742
ur
Herndl, G.J., Reinthaler, T., 2013. Microbial control of the dark end of the biological pump. Nat.
Jo
Geosci. 6, 718–724. https://doi.org/10.1038/ngeo1921 Herrle, J.O., Kößler, P., Friedrich, O., Erlenkeuser, H., Hemleben, C., 2004. High-resolution carbon isotope records of the Aptian to Lower Albian from SE France and the Mazagan Plateau (DSDP Site 545): a stratigraphic tool for paleoceanographic and paleobiologic reconstruction. Earth Planet. Sci. Lett. 218, 149–161. https://doi.org/10.1016/S0012821X(03)00646-0 Hesselbo, S.P., Hudson, A.J.L., Huggett, J.M., Leng, M.J., Riding, J.B., Ullmann, C. V., 2019. Palynological, geochemical, and mineralogical characteristics of the Early Jurassic
Journal Pre-proof Liasidium Event in the Cleveland Basin, Yorkshire, UK. Newsletters Stratigr. https://doi.org/10.1127/nos/2019/0536 Hesselbo, S.P., Jenkyns, H.C., Duarte, L. V., Oliveira, L.C. V, 2007. Carbon-isotope record of the Early Jurassic (Toarcian) Oceanic Anoxic Event from fossil wood and marine carbonate (Lusitanian Basin, Portugal). Earth Planet. Sci. Lett. 253, 455–470. https://doi.org/10.1016/j.epsl.2006.11.009
of
Hesselbo, S.P., Meister, C., Gröcke, D.R., 2000. A potential global stratotype for the
ro
Sinemurian–Pliensbachian boundary (Lower Jurassic), Robin Hood’s Bay, UK: ammonite
-p
faunas and isotope stratigraphy. Geol. Mag. 137, 601–607.
re
https://doi.org/10.1017/S0016756800004672
Hesselbo, S.P., Morgans‐ Bell, H.S., McElwain, J.C., Rees, P.M., Robinson, S.A., Ross, C.E.,
lP
2003. Carbon‐ Cycle Perturbation in the Middle Jurassic and Accompanying Changes in
na
the Terrestrial Paleoenvironment. J. Geol. 111, 259–276. https://doi.org/10.1086/373968 Hoefs, J., 2015. Stable Isotope Geochemistry. Springer International Publishing.
ur
https://doi.org/10.1007/978-3-319-19716-6
Jo
Holmden, C., Creaser, R.A., Muehlenbachs, K., Leslie, S.A., Bergström, S.M., 1998. Isotopic evidence for geochemical decoupling between ancient epeiric seas and bordering oceans: implications for secular curves. Geology 26, 567–570. https://doi.org/10.1130/00917613(1998)026<0567:IEFGDB>2.3.CO;2 Hubbard, R.N., Boulter, M.C., 1997. Mid Mesozoic floras and climates. Palaeontology 40, 43– 70. Ikeda, M., Bôle, M., Baumgartner, P.O., 2016. Orbital-scale changes in redox condition and biogenic silica/detrital fluxes of the Middle Jurassic Radiolarite in Tethys (Sogno,
Journal Pre-proof Lombardy, N-Italy): Possible link with glaciation? Palaeogeogr. Palaeoclimatol. Palaeoecol. 457, 247–257. https://doi.org/10.1016/j.palaeo.2016.06.009 Imlay, R.W., 1981. Jurassic (Bathonian and Callovian) Ammonites in Eastern Oregon and Western Idaho, Geological Survey professional paper. U.S. Dept. of the Interior, Geological Survey, Washington, D.C. Immenhauser, A., Della Porta, G., Kenter, J. a M., Bahamonde, J.R., 2003. An alternative model
of
for positive shifts in shallow-marine carbonate δ 13 C and δ 18 O. Sedimentology 50, 953–
ro
959. https://doi.org/10.1046/j.1365-3091.2003.00590.x
-p
Jach, R., Djerić, N., Goričan, Š., Reháková, D., 2014. Integrated Stratigraphy of The Middle–
re
Upper Jurassic of the Krížna Nappe, Tatra Mountains. Ann. Soc. Geol. Pol. 84, 1–33. Jacobsen, S.B., Kaufman, A.J., 1999. The Sr, C and O isotopic evolution of Neoproterozoic
lP
seawater. Chem. Geol. 161, 37–57. https://doi.org/10.1016/S0009-2541(99)00080-7
na
Jarvis, I., Trabucho-Alexandre, J., Gröcke, D.R., Uličný, D., Laurin, J., 2015. Intercontinental correlation of organic carbon and carbonate stable isotope records: evidence of climate and
ur
sea-level change during the Turonian (Cretaceous). Depos. Rec. 1, 53–90.
Jo
https://doi.org/10.1002/dep2.6
Jenkyns, H.C., 1996. Relative sea-level change and carbon isotopes: data from the Upper Jurassic (Oxfordian) of central and Southern Europe. Terra Nov. 8, 75–85. https://doi.org/10.1111/j.1365-3121.1996.tb00727.x Jenkyns, H.C., Clayton, C.J., 1986. Black shales and carbon isotopes in pelagic sediments from the Tethyan Lower Jurassic. Sedimentology 33, 87–106. https://doi.org/10.1111/j.13653091.1986.tb00746.x Jenkyns, H.C., Jones, C., Gröcke, D., Hesselbo, S., Parkinson, D., 2002. Chemostratigraphy of
Journal Pre-proof the Jurassic System: applications, limitations and implications for palaeoceanography. J. Geol. Soc. London. 159, 351–378. https://doi.org/10.1144/0016-764901-130 Jenkyns, H.C., Schouten-Huibers, L., Schouten, S., Sinninghe Damsté, J.S., 2012. Warm Middle Jurassic–Early Cretaceous high-latitude sea-surface temperatures from the Southern Ocean. Clim. Past 8, 215–226. https://doi.org/10.5194/cp-8-215-2012 Jiao, N., Herndl, G.J., Hansell, D.A., Benner, R., Kattner, G., Wilhelm, S.W., Kirchman, D.L.,
of
Weinbauer, M.G., Luo, T., Chen, F., Azam, F., 2010. Microbial production of recalcitrant
-p
593–599. https://doi.org/10.1038/nrmicro2386
ro
dissolved organic matter: long-term carbon storage in the global ocean. Nat Rev Micro 8,
re
Jiao, N., Robinson, C., Azam, F., Thomas, H., Baltar, F., Dang, H., Hardman-Mountford, N.J., Johnson, M., Kirchman, D.L., Koch, B.P., Legendre, L., Li, C., Liu, J., Luo, T., Luo, Y.-W.,
lP
Mitra, A., Romanou, A., Tang, K., Wang, X., Zhang, C., Zhang, R., 2014. Mechanisms of
na
microbial carbon sequestration in the ocean – future research directions. Biogeosciences 11, 5285–5306. https://doi.org/10.5194/bg-11-5285-2014
ur
Johnston, D.T., Macdonald, F.A., Gill, B.C., Hoffman, P.F., Schrag, D.P., 2012. Uncovering the
Jo
Neoproterozoic carbon cycle. Nature 483, 320–323. https://doi.org/10.1038/nature10854 Kenig, F.G., Hayes, J.M., Popp, B.N., Summons, R.E., 1994. Isotopic biogeochemistry of the Oxford Clay Formation (Jurassic), UK. J. Geol. Soc. London. 151, 139–152. https://doi.org/10.1144/gsjgs.151.1.0139 Koevoets, M.J., Abay, T.B., Hammer, Ø., Olaussen, S., 2016. High-resolution organic carbon– isotope stratigraphy of the Middle Jurassic–Lower Cretaceous Agardhfjellet Formation of central Spitsbergen, Svalbard. Palaeogeogr. Palaeoclimatol. Palaeoecol. 449, 266–274. https://doi.org/10.1016/j.palaeo.2016.02.029
Journal Pre-proof Krissansen-Totton, J., Buick, R., Catling, D.C., 2015. A statistical analysis of the carbon isotope record from the Archean to Phanerozoic and implications for the rise of oxygen. Am. J. Sci. 315, 275–316. https://doi.org/10.2475/04.2015.01 Kump, L.R., Arthur, M.A., 1999. Interpreting carbon-isotope excursions: carbonates and organic matter. Chem. Geol. 161, 181–198. https://doi.org/10.1016/S0009-2541(99)00086-8 Kusuhashi, N., Matsuoka, H., Kamiya, H., Setoguchi, T., 2002. Stratigraphy of the late Mesozoic
of
Tetori Group in the Hakusan Region, central Japan : an overview. Mem. Fac. Sci. Kyoto
ro
Univ. Ser. Geol. Mineral. 59, 9–31.
-p
Legendre, L., Rivkin, R.B., Weinbauer, M.G., Guidi, L., Uitz, J., 2015. The microbial carbon
re
pump concept: Potential biogeochemical significance in the globally changing ocean. Prog. Oceanogr. 134, 432–450. https://doi.org/10.1016/j.pocean.2015.01.008
lP
Lewan, M.D., 1983. Effects of thermal maturation on stable organic carbon isotopes as
na
determined by hydrous pyrolysis of Woodford Shale. Geochim. Cosmochim. Acta 47, 1471–1479. https://doi.org/10.1016/0016-7037(83)90306-X
ur
Mangold, C., 1990. Le Bathonien du Cap Mondego (N de Figueira da Foz, Portugal):
Jo
Biochronologie et Correlations. Cah. Univ. Catho Lyon, Sér. Sci. 4, 89–105. Marshall, J.D., 1992. Climatic and oceanographic isotopic signals from the carbonate rock record and their preservation. Geol. Mag. 129, 143. https://doi.org/10.1017/S0016756800008244 Mitchell, S.F., Paul, C.R.C., Gale, A.S., 1996. Carbon isotopes and sequence stratigraphy. Geol. Soc. London, Spec. Publ. 104, 11–24. Morton, N., Hudson, J.D., 1995. Field guide to the Jurassic of the Isles of Raasay and Skye, Inner Hebrides, NW Scotland, in: Taylor, P.D. (Ed.), Field Geology of the British Jurassic. The Geological Society, London, pp. 209–280.
Journal Pre-proof Mukhopadhyay, P.K., Wade, J.A., 1990. Organic facies and maturation of sediments from three Scotian Shelf wells. Bull. Can. Pet. Geol. 38, 407–425. Nagy, J., Finstad, E.K., Dypvik, H., Merethe, A., Bremer, G.A., Bremer, M.G.A., 2001. Response of foraminiferal facies to Transgressive-Regressive cycles in the Callovian of northeast Scotland. J. Foraminifer. Res. 31, 324–349. https://doi.org/10.2113/0310324 Nebbioso, A., Piccolo, A., 2013. Molecular characterization of dissolved organic matter (DOM):
of
a critical review. Anal. Bioanal. Chem. 405, 109–124. https://doi.org/10.1007/s00216-012-
ro
6363-2
-p
Newton, R., Bottrell, S., 2007. Stable isotopes of carbon and sulphur as indicators of
re
environmental change: past and present. J. Geol. Soc. London. 164, 691–708. Nunn, E. V, Price, G.D., Hart, M.B., Page, K.N., Leng, M.J., 2009. Isotopic signals from
lP
Callovian–Kimmeridgian (Middle–Upper Jurassic) belemnites and bulk organic carbon,
na
Staffin Bay, Isle of Skye, Scotland. J. Geol. Soc. London. 166, 633. https://doi.org/https://doi.org/10.1144/0016-76492008-067
ur
O’Dogherty, L., Sandoval, J., Bartolini, A., Bruchez, S., Bill, M., Guex, J., 2006. Carbon–isotope
Jo
stratigraphy and ammonite faunal turnover for the Middle Jurassic in the Southern Iberian palaeomargin. Palaeogeogr. Palaeoclimatol. Palaeoecol. 239, 311–333. https://doi.org/10.1016/j.palaeo.2006.01.018 Oehlert, A.M., Swart, P.K., 2014. Interpreting carbonate and organic carbon isotope covariance in the sedimentary record. Nat. Commun. 5, 4672. https://doi.org/10.1038/ncomms5672 Ogawa, H., Tanoue, E., 2003. Dissolved organic matter in oceanic waters. J. Oceanogr. 59, 129– 147. https://doi.org/10.1023/A:1025528919771 Ogg, J.G., Ogg, G.M., Gradstein, F.M., Ogg, J.G., Ogg, G.M., Gradstein, F.M., 2016. 12 –
Journal Pre-proof Jurassic, in: A Concise Geologic Time Scale. Elsevier, pp. 151–166. https://doi.org/10.1016/B978-0-444-59467-9.00012-1 Pahlow, M., Riebesell, U., 2000. Temporal Trends in Deep Ocean Redfield Ratios. Science (80-. ). 287, 831–833. https://doi.org/10.1126/science.287.5454.831 Panchuk, K.M., Holmden, C.E., Leslie, S.A., 2006. Local Controls on Carbon Cycling in the Ordovician Midcontinent Region of North America, with Implications for Carbon Isotope
of
Secular Curves. J. Sediment. Res. 76, 200–211. https://doi.org/10.2110/jsr.2006.017
ro
Patterson, W.P., Walter, L.M., 1994. Depletion of 13C in seawater ΣCO2 on modern carbonate
-p
platforms : Significance for the carbon isotopic record of carbonates. Geology 22, 885–888.
re
https://doi.org/10.1130/0091-7613(1994)022<0885
Pavia, G., Enay, R., 1997. Definition of the Aalenian-Bajocian Stage boundary. Episodes 20, 16–
lP
22.
na
Peltier, W.R., Liu, Y., Crowley, J.W., 2007. Snowball Earth prevention by dissolved organic carbon remineralization. Nature 450, 813–818. https://doi.org/10.1038/nature06354
ur
Peters, K.E., 1986. Guidelines for evaluating petroleum source rock using programmed
Jo
pyrolysis. Am. Assoc. Pet. Geol. Bull. 70, 318–329. Pinheiro, L.M., Wilson, R.C.L., Pena dos Reis, R., Whitmarsh, R.B., Ribeiro, A., 1996. The western Iberia Margin: a geophysical and geological overview, in: Whitmarsh, R.B., Sawyer, D.S., Klaus, A., Masson, D.G. (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results, Vol 149. Ocean Drilling Program, pp. 3–23. https://doi.org/10.2973/odp.proc.sr.149.246.1996 Polimene, L., Sailley, S., Clark, D., Mitra, A., Allen, J.I., 2017. Biological or microbial carbon pump? The role of phytoplankton stoichiometry in ocean carbon sequestration. J. Plankton
Journal Pre-proof Res. 39, 180–186. https://doi.org/10.1093/plankt/fbw091 Price, G.D., 1999. The evidence and implications of polar ice during the Mesozoic. EarthScience Rev. 48, 183–210. https://doi.org/10.1016/S0012-8252(99)00048-3 Rahmstorf, S., Box, J.E., Feulner, G., Mann, M.E., Robinson, A., Rutherford, S., Schaffernicht, E.J., 2015. Exceptional twentieth-century slowdown in Atlantic Ocean overturning circulation. Nat. Clim. Chang. 5, 475–480. https://doi.org/10.1038/nclimate2554
of
Rasmussen, E.S., Lomholt, S., Andersen, C., Vejbæk, O. V., 1998. Aspects of the structural
ro
evolution of the Lusitanian Basin in Portugal and the shelf and slope area offshore Portugal.
-p
Tectonophysics 300, 199–225. https://doi.org/10.1016/S0040-1951(98)00241-8
re
Ridgwell, A., Arndt, S., 2015. Why Dissolved Organics Matter: DOC in Ancient Oceans and Past Climate Change, in: Hansell, D.A., Carlson, C.A. (Eds.), Biogeochemistry of Marine
lP
Dissolved Organic Matter. Elsevier, pp. 1–20. https://doi.org/10.1016/B978-0-12-405940-
na
5.00001-7
Rothman, D.H., Hayes, J.M., Summons, R.E., 2003. Dynamics of the Neoproterozoic carbon
ur
cycle. Proc. Natl. Acad. Sci. 100, 8124–8129. https://doi.org/10.1073/pnas.0832439100
Jo
Ruget-Perrot, C., 1961. Etudes stratigraphiques sur le Dogger et le Maim inferieur du Portugal au Nord du Tage. Memorias dos Serviços Geológicos Port. 7, 1–197. Rush, D., Talbot, H.M., van der Meer, M.T.J., Hopmans, E.C., Douglas, B., Sinninghe Damsté, J.S., 2019. Biomarker evidence for the occurrence of anaerobic ammonium oxidation in the eastern Mediterranean Sea during Quaternary and Pliocene sapropel formation. Biogeosciences 16, 2467–2479. https://doi.org/10.5194/bg-16-2467-2019 Schidlowski, M., 1988. A 3,800-million-year isotopic record of life from carbon in sedimentary rocks. Nature 333, 313–318. https://doi.org/10.1038/333313a0
Journal Pre-proof Schidlowski, M., Appel, P.W.U., Eichmann, R., Junge, C.E., 1979. Carbon isotope geochemistry of the 3.7 × 109-yr-old Isua sediments, West Greenland: implications for the Archaean carbon and oxygen cycles. Geochim. Cosmochim. Acta 43, 189–199. https://doi.org/10.1016/0016-7037(79)90238-2 Schlanger, S., Jenkyns, H., 1976. Cretaceous Oceanic Anoxic Events: Causes and Consequences. Geol. en Mijnb. 55, 179–184.
of
Schmittner, A., Oschlies, A., Matthews, H.D., Galbraith, E.D., 2008. Future changes in climate,
ro
ocean circulation, ecosystems, and biogeochemical cycling simulated for a business-as-
re
https://doi.org/10.1029/2007GB002953
-p
usual CO 2 emission scenario until year 4000 AD. Global Biogeochem. Cycles 22, n/a-n/a.
Sexton, P.F., Norris, R.D., Wilson, P.A., Palike, H., Westerhold, T., Rohl, U., Bolton, C.T.,
lP
Gibbs, S., 2011. Eocene global warming events driven by ventilation of oceanic dissolved
na
organic carbon. Nature 471, 349–352. https://doi.org/10.1038/nature09826 Shen, Y., Benner, R., 2018. Mixing it up in the ocean carbon cycle and the removal of refractory
ur
dissolved organic carbon. Sci. Rep. 8, 2542. https://doi.org/10.1038/s41598-018-20857-5
Jo
Silva, R.L., Duarte, L.V., 2015. Organic matter production and preservation in the Lusitanian Basin (Portugal) and Pliensbachian climatic hot snaps. Glob. Planet. Change 131, 24–34. https://doi.org/10.1016/j.gloplacha.2015.05.002 Silva, R.L., Duarte, L.V., Comas-Rengifo, M., 2015. Facies and Carbon Isotope Chemostratigraphy of Lower Jurassic Carbonate Deposits, Lusitanian Basin (Portugal): Implications and Limitations to the Application in Sequence Stratigraphic Studies, Chemostratigraphy: Concepts, Techniques, and Applications. https://doi.org/10.1016/B9780-12-419968-2.00013-3
Journal Pre-proof Silva, R.L., Duarte, L.V., Mendonça Filho, J.G., 2013. Optical and geochemical characterization of Upper Sinemurian (Lower Jurassic) fossil wood from the Lusitanian Basin (Portugal). Geochem. J. 47, 489–498. https://doi.org/10.2343/geochemj.2.0270 Silva, R.L., Duarte, L. V., Comas-Rengifo, M.J., Mendonça Filho, J.G., Azerêdo, A.C., 2011. Update of the carbon and oxygen isotopic records of the Early-Late Pliensbachian (Early Jurassic, ~187Ma): Insights from the organic-rich hemipelagic series of the Lusitanian
of
Basin (Portugal). Chem. Geol. 283, 177–184.
ro
https://doi.org/10.1016/j.chemgeo.2011.01.010
-p
Soares, A.F., Duarte, L.V., 1997. Tectonic and eustatic signatures in the Lower and Middle
re
Jurassic of the Lusitanian Basin, in: IV Congreso Del Jurasico de Espana. Terruel, pp. 111– 114.
lP
Soares, A.F., Rocha, R.B., Elmi, S., Henriques, M.H., Mouterde, R., Almeras, Y., Ruget, C.,
na
Marques, J., Duarte, L., Carapito, M.C., Kullberg, J.C., 1993. The North-Lusitanian Subbasin (Portugal) from the Triassic to the Middle Jurassic: a model of an “aborted rift.”
ur
Comptes rendus l’Académie des Sci. 317, 1659–1666.
Jo
Song, Haijun, Jiang, G., Poulton, S.W., Wignall, P.B., Tong, J., Song, Huyue, An, Z., Chu, D., Tian, L., She, Z., Wang, C., 2017. The onset of widespread marine red beds and the evolution of ferruginous oceans. Nat. Commun. 8, 399. https://doi.org/10.1038/s41467-01700502-x Suan, G., van de Schootbrugge, B., Adatte, T., Fiebig, J., Oschmann, W., 2015. Calibrating the magnitude of the Toarcian carbon cycle perturbation. Paleoceanography 30, 495–509. https://doi.org/10.1002/2014PA002758 Sundquist, E.T., 1991. Steady- and non-steady-state carbonate-silicate controls on atmospheric
Journal Pre-proof CO2. Quat. Sci. Rev. 10, 283–296. https://doi.org/10.1016/0277-3791(91)90026-Q Sundquist, Eric T., 1985. Geological Perspectives on Carbon Dioxide and the Carbon Cycle, in: Sundquist, E.T., Broecker, W.S. (Eds.), The Carbon Cycle and Atmospheric CO2: Natural Variations Archean to Present, Vol 32. pp. 55–59. https://doi.org/10.1029/GM032p0005 Swanson-Hysell, N.L., Rose, C. V, Calmet, C.C., Halverson, G.P., Hurtgen, M.T., Maloof, A.C., 2010. Cryogenian Glaciation and the Onset of Carbon-Isotope Decoupling. Science (80-. ).
of
328, 608–611. https://doi.org/10.1126/science.1184508
ro
Swart, P.K., 2015. The geochemistry of carbonate diagenesis: The past, present and future.
-p
Sedimentology 62, 1233–1304. https://doi.org/10.1111/sed.12205
re
Swart, P.K., Eberli, G., 2005. The nature of the δ13C of periplatform sediments: Implications for stratigraphy and the global carbon cycle. Sediment. Geol. 175, 115–129.
lP
https://doi.org/10.1016/j.sedgeo.2004.12.029
na
Swart, P.K., Melim, L.A., 2000. The origin of dolomites in Tertiary sediments from the Margin of Great Bahama Bank. J. Sediment. Res. 70, 738–748.
ur
Thibault, N., Ruhl, M., Ullmann, C. V., Korte, C., Kemp, D.B., Gröcke, D.R., Hesselbo, S.P.,
Jo
2018. The wider context of the Lower Jurassic Toarcian oceanic anoxic event in Yorkshire coastal outcrops, UK. Proc. Geol. Assoc. 129, 372–391. https://doi.org/10.1016/j.pgeola.2017.10.007 Tyson, R. V., 1995. Sedimentary organic matter: organic facies and palynofacies. Springer Netherlands. Verdugo, P., Alldredge, A.L., Azam, F., Kirchman, D.L., Passow, U., Santschi, P.H., 2004. The oceanic gel phase: a bridge in the DOM–POM continuum. Mar. Chem. 92, 67–85. https://doi.org/10.1016/j.marchem.2004.06.017
Journal Pre-proof Wallmann, K., Aloisi, G., 2012. The Global Carbon Cycle: Geological Processes, in: Knoll, A.H., Canfield, D.E., Konhauser, K.O. (Eds.), Fundamentals of Geobiology. John Wiley & Sons, Ltd, Chichester, UK, pp. 20–35. https://doi.org/10.1002/9781118280874.ch3 Wefer, G., Berger, W.H., 1991. Isotope paleontology: growth and composition of extant calcareous species. Mar. Geol. 100, 207–248. https://doi.org/10.1016/0025-3227(91)90234U
of
Weissert, H., Lini, A., Föllmi, K.B., Kuhn, O., 1998. Correlation of Early Cretaceous carbon
ro
isotope stratigraphy and platform drowning events: A possible link? Palaeogeogr.
-p
Palaeoclimatol. Palaeoecol. 137, 189–203. https://doi.org/10.1016/S0031-0182(97)00109-0
re
Wilson, J.D., Arndt, S., 2017. Modeling radiocarbon constraints on the dilution of dissolved organic carbon in the deep ocean. Global Biogeochem. Cycles 31, 775–786.
lP
https://doi.org/10.1002/2016GB005520
na
Wilson, R.C.L., Hiscott, R.N., Willis, M.G., Gradstein, F.M., 1989. The Lusitanian Basin of West-Central Portugal: Mesozoic and Tertiary Tectonic, Stratigraphic, and Subsidence
ur
History: Chapter 22: European-African Margin. Extensional Tectonics Stratigr. North Atl.
Jo
Margins 46, 341–361.
Wohlers, J., Engel, A., Zöllner, E., Breithaupt, P., Jürgens, K., Hoppe, H.-G., Sommer, U., Riebesell, U., 2009. Changes in biogenic carbon flow in response to sea surface warming. Proc. Natl. Acad. Sci. 106, 7067–7072. https://doi.org/10.1073/pnas.0812743106 Yager, J.A., West, A.J., Corsetti, F.A., Berelson, W.M., Rollins, N.E., Rosas, S., Bottjer, D.J., 2017. Duration of and decoupling between carbon isotope excursions during the endTriassic mass extinction and Central Atlantic Magmatic Province emplacement. Earth Planet. Sci. Lett. 473, 227–236. https://doi.org/10.1016/j.epsl.2017.05.031
Journal Pre-proof Yin, J., 2007. A review on Jurassic sea-level changes in Himalayan Tibet. Beringeria 37, 253–
Jo
ur
na
lP
re
-p
ro
of
266.
Journal Pre-proof Figure captions Figure 1. a) Cartoon illustrating the main oceanic carbon reservoirs and selected transfer processes (Emerson and Hedges, 2008; Jiao et al., 2010, 2014; Hansell and Carlson, 2013; Carlson and Hansell, 2015; Ridgwell and Arndt, 2015); b) Schematic depiction of the sizes of aggregates containing organic carbon in the ocean (modified from Verdugo et al., 2004; Emerson and Hedges, 2008). 1 - air-water gas exchange; 2 - respiration; 3 - primary production
of
(photosynthesis); 4 - resuspension/erosion/influx/oxidation; 5 -microbial biomass (heterotrophy);
ro
6 - phytoplankton production, exudation, grazing, viral lysis, solubilization from POC, riverine
-p
input; 7 - respiration; 8 - respiration/oxidation; 9 - dissociation; 10 - carbonate production and precipitation; 11 - carbonate dissolution; 12 - dissolution; 13 - sedimentation; 14 – microbial
re
activity; 15 – phototransformation; 16 - sedimentation (PIC)/precipitation (DIC); 17 –
na
lP
methanogenesis.
Figure 2. Simplified geological map of the central-northern part of the Lusitanian Basin
Jo
ur
(Portugal) and distribution of the carbonate units of Lower Jurassic age.
Figure 3. Stratigraphic units and main sedimentological features of the Lower and Middle Jurassic series in the Lusitanian Basin (Portugal) (modified from (Azerêdo et al., 2020). Major Transgressive-Regressive and Transgressive-Regressive facies cycles from (Azerêdo et al., 2014). Sedimentary discontinuities from Soares and Duarte (1997).
Figure 4. Stratigraphic log, TOC, and kerogen and carbonate δ13C record of the Late Bathonian– Lower Callovian boundary in the Cabo Mondego Formation, Lusitanian Basin (Portugal).
Journal Pre-proof
Figure 5. δ13C versus δ18O cross plot of the analysed samples from the Late Bathonian–Lower Callovian boundary in the Cabo Mondego Formation, Lusitanian Basin (Portugal).
Figure 6. Stratigraphic log, TOC and carbonate δ13C record of the Middle Bathonian–Lower Callovian boundary in the Cabo Mondego Formation, Lusitanian Basin (Portugal) and
of
comparison with the δ13C record of several other European basins; Umbria-Marche Basin
ro
(Bartolini et al., 1996); Southern Alpes (Jenkyns, 1996), External Subbetic Basin (O’Dogherty et
re
-p
al., 2006) (LOESS smoothing, smoothing factor = 0.2, confidence interval = 95%).
Figure 7. Cartoon illustrating the relative sizes and variation of the several carbon pools in the
lP
oceans in the Bathonian–Callovian. The relative size, chemical and isotopic composition, and
na
oceanographical distribution of the DPOOC, DIC, and atmospheric CO2 pools and composition and magnitude of the fluxes between the different superficial carbon reservoirs, including
Jo
ur
external inputs, are unknown.
Journal Pre-proof Highlights
ur
na
lP
re
-p
ro
of
At least one δ13Ccarb-TOC decoupling event occurred in the Late Bathonian–Callovian Decoupling resulted from the accumulation of oceanic organic carbon (OOC) The enlargement of the OOC pool increased C residence time and buffered δ13CTOC to change The wane and wax of the OOC pool resulted in carbon isotopic excursions These decoupling events are evidence a non-steady state carbon cycle
Jo
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7