Hostname: page-component-76fb5796d-wq484 Total loading time: 0 Render date: 2024-04-25T09:42:38.889Z Has data issue: false hasContentIssue false

The response of water column and sedimentary environments to the advent of the Messinian salinity crisis: insights from an onshore deep-water section (Govone, NW Italy)

Published online by Cambridge University Press:  07 September 2020

Mathia Sabino
Affiliation:
Institut für Geologie, Centrum für Erdsystemforschung und Nachhaltigkeit, Universität Hamburg, D-20146Hamburg, Germany
Francesco Dela Pierre
Affiliation:
Dipartimento di Scienze della Terra, Università degli Studi di Torino, I-10125Torino, Italy
Marcello Natalicchio
Affiliation:
Dipartimento di Scienze della Terra, Università degli Studi di Torino, I-10125Torino, Italy
Daniel Birgel
Affiliation:
Institut für Geologie, Centrum für Erdsystemforschung und Nachhaltigkeit, Universität Hamburg, D-20146Hamburg, Germany
Susanne Gier
Affiliation:
Department für Geodynamik und Sedimentologie, Universität Wien, A-1090Wien, Austria
Jörn Peckmann*
Affiliation:
Institut für Geologie, Centrum für Erdsystemforschung und Nachhaltigkeit, Universität Hamburg, D-20146Hamburg, Germany
*
Author for correspondence: Jörn Peckmann, Email: joern.peckmann@uni-hamburg.de
Rights & Permissions [Opens in a new window]

Abstract

During Messinian time, the Mediterranean underwent hydrological modifications culminating 5.97 Ma ago with the Messinian salinity crisis (MSC). Evaporite deposition and alleged annihilation of most marine eukaryotes were taken as evidence of the establishment of basin-wide hypersalinity followed by desiccation. However, the palaeoenvironmental conditions during the MSC are still a matter of debate, chiefly because most of its sedimentary record is buried below the abyssal plains of the present-day Mediterranean Sea. To shed light on environmental change at the advent and during the early phase of the MSC, we investigated the Govone section from the Piedmont Basin (NW Italy) using a multidisciplinary approach (organic geochemical, petrographic, and carbon and oxygen stable isotope analyses). The Govone section archives the onset of the crisis in a succession of organic-rich shales and dolomite-rich marls. The MSC part of the succession represents the deep-water equivalent of sulphate evaporites deposited at the basin margins during the first phase of the crisis. Our study reveals that the onset of the MSC was marked by the intensification of water-column stratification, rather than the establishment of widespread hypersaline conditions. A chemocline divided the water column into an oxygen-depleted, denser and more saline bottom layer and an oxygenated, upper seawater layer influenced by freshwater inflow. Vertical oscillations of the chemocline controlled the stratigraphic architecture of the sediments pertaining to the first stage of the MSC. Accordingly, temporal and spatial changes of water masses with different redox chemistries must be considered when interpreting the MSC event.

Type
Original Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2020. Published by Cambridge University Press

1. Introduction

During Messinian time (7.25–5.33 Ma) the Mediterranean area underwent progressive environmental and hydrological changes, which culminated c. 5.97 Ma ago with the Messinian salinity crisis (MSC; Hsü et al. Reference Hsü, Ryan and Cita1973; Roveri et al. Reference Roveri, Flecker, Krijgsman, Lofi, Lugli, Manzi, Sierro, Bertini, Camerlenghi, De Lange, Govers, Hilgen, Hübscher, Meijer and Stoica2014; Camerlenghi & Aloisi, Reference Camerlenghi and Aloisi2020). During this event, the Mediterranean Sea turned into the youngest Salt Giant of Earth history (Warren, Reference Warren2010) as a response of its isolation from the global ocean caused by the restriction of the Mediterranean–Atlantic gateways (Flecker et al. Reference Flecker, Krijgsman, Capella, de Castro Martíns, Dmitrieva, Mayser, Marzocchi, Modestu, Ochoa, Simon, Tulbure, van den Berg, van der Schee, de Lange, Ellam, Govers, Gutjahr, Hilgen, Kouwenhoven, Lofi, Meijer, Sierro, Bachiri, Barhoun, Alami, Chacon, Flores, Gregory, Howard, Lunt, Ochoa, Pancost, Vincent and Yousfi2015; Krijgsman et al. Reference Krijgsman, Capella, Simon, Hilgen, Kouwenhoven, Meijer, Sierro, Tulbure, van den Berg, van der Schee and Flecker2018; Capella et al. Reference Capella, Flecker, Hernández-Molina, Simon, Meijer, Rogerson, Sierro and Krijgsman2019). One of the most striking effects of the hydrological changes that affected the basin was the intensification of water-column stratification and consequent development of bottom-water anoxia starting at c. 6.7 Ma (Roveri et al. Reference Roveri, Flecker, Krijgsman, Lofi, Lugli, Manzi, Sierro, Bertini, Camerlenghi, De Lange, Govers, Hilgen, Hübscher, Meijer and Stoica2014). The deterioration of palaeoenvironmental conditions was recorded by the deposition of precession-paced alternations of organic-rich shales, diatomites and marls (Hilgen & Krijgsman, Reference Hilgen and Krijgsman1999; Sierro et al. Reference Sierro, Flores, Zamarreño, Vázquez, Utrilla, Francés, Hilgen and Krijgsman1999; Krijgsman et al. Reference Krijgsman, Blanc-Valleron, Flecker, Hilgen, Kouwenhoven, Merle, Orszag-Sperber and Rouchy2002). Starting from 5.97 Ma (Manzi et al. Reference Manzi, Gennari, Hilgen, Krijgsman, Lugli, Roveri and Sierro2013), the shale-diatomite-marl successions were replaced at the basin margins by couplets of sulphate evaporites and organic-rich shales of the Primary Lower Gypsum unit (PLG; Roveri et al. Reference Roveri, Lugli, Manzi and Schreiber2008), marking the first stage of the MSC (5.97–5.60 Ma; Roveri et al. Reference Roveri, Flecker, Krijgsman, Lofi, Lugli, Manzi, Sierro, Bertini, Camerlenghi, De Lange, Govers, Hilgen, Hübscher, Meijer and Stoica2014). The deposition of evaporites has been considered as compelling evidence of the development of hypersaline conditions, which resulted in the demise of most marine eukaryotes (e.g. Bellanca et al. Reference Bellanca, Caruso, Ferruzza, Neri, Rouchy, Sprovieri and Blanc-Valleron2001), but promoted the rise of halophilic prokaryotes (Turich & Freeman, Reference Turich and Freeman2011; Birgel et al. Reference Birgel, Guido, Liu, Hinrichs, Gier and Peckmann2014). The PLG unit passes laterally in intermediate- to deep-water settings (> 200 m depth) into evaporite-free successions, composed of shales alternating with carbonates and/or dolomite-rich marls barren of calcareous microfossils (Manzi et al. Reference Manzi, Roveri, Gennari, Bertini, Biffi, Giunta, Iaccarino, Lanci, Lugli, Negri, Riva, Rossi and Taviani2007, Reference Manzi, Lugli, Roveri, Schreiber and Gennari2011; Dela Pierre et al. Reference Dela Pierre, Bernardi, Cavagna, Clari, Gennari, Irace, Lozar, Lugli, Manzi, Natalicchio, Roveri and Violanti2011; Roveri et al. Reference Roveri, Flecker, Krijgsman, Lofi, Lugli, Manzi, Sierro, Bertini, Camerlenghi, De Lange, Govers, Hilgen, Hübscher, Meijer and Stoica2014; Natalicchio et al. Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019).

After the formulation of the ‘deep desiccated basin model’ in the 1970s (Hsü et al. Reference Hsü, Ryan and Cita1973), the palaeoenvironmental conditions of the Mediterranean water column and seafloor during the MSC are still a matter of debate. The scarcity or lack of body fossils in MSC sediments, the lack of modern analogues for the Messinian evaporites and the inaccessibility of the offshore evaporites have hampered the development of a widely accepted scenario (Rouchy & Caruso, Reference Rouchy and Caruso2006; Ryan, Reference Ryan2009; Roveri et al. Reference Roveri, Flecker, Krijgsman, Lofi, Lugli, Manzi, Sierro, Bertini, Camerlenghi, De Lange, Govers, Hilgen, Hübscher, Meijer and Stoica2014; Camerlenghi & Aloisi, Reference Camerlenghi and Aloisi2020).

The analysis of molecular fossils (lipid biomarkers) preserved in Messinian evaporites and their intermediate- to deep-water lateral equivalents provides fundamental information for the reconstruction of the Mediterranean hydrological cycle and palaeoenvironments (Vasiliev et al. Reference Vasiliev, Mezger, Lugli, Reichart, Manzi and Roveri2017; Natalicchio et al. Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020). This approach is a valuable tool to unveil the conditions in the water column during the MSC. Organic geochemical investigations targeting abyssal evaporites (Christeleit et al. Reference Christeleit, Brandon and Zhuang2015) and Messinian salts (Reference Isaji, Yoshimura, Kuroda, Tamenori, Jiménez-Espejo, Lugli, Manzi, Roveri, Kawahata and OhkouchiIsaji et al. 2019b ) revealed that evaporites were deposited under a stratified water column during the latest MSC phases. These results agree with recent hydrological models (de Lange & Krijgsman, Reference de Lange and Krijgsman2010; Simon & Meijer, Reference Simon and Meijer2017; García-Veigas et al. Reference García-Veigas, Cendón, Gibert, Lowenstein and Artiaga2018). In contrast, only a few geochemical data are available for the sediments deposited during the early stages of the MSC (Kenig et al. Reference Kenig, Sinninghe Damsté, Frewin, Hayes and De Leeuw1995; Reference Sinninghe Damsté, Kenig, Koopmans, Köster, Schouten, Hayes and de LeeuwSinninghe Damsté et al. 1995b ; Reference Isaji, Kawahata, Takano, Ogawa, Kuroda, Yoshimura, Lugli, Manzi, Roveri and OhkouchiIsaji et al. 2019a ). These studies all describe organic-rich shales of the PLG unit, depicting a stratified basin that received freshwater from rivers (Natalicchio et al. Reference Natalicchio, Birgel, Peckmann, Lozar, Carnevale, Liu, Hinrichs and Dela Pierre2017, Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020) and/or low-salinity water from the Paratethys (Grothe et al. Reference Grothe, Andreetto, Reichart, Wolthers, Van Baak, Vasiliev, Stoica, Sangiorgi, Middelburg, Davies and Krijgsman2020).

To shed new light on the response of the water column and sediments to the advent of the MSC, we studied sedimentary strata exposed in the Govone section (Piedmont Basin, NW Italy). In this section, the onset of the crisis is archived in a sequence of organic-rich shales and marls, representing the deep-water equivalents of primary sulphate evaporites deposited at the basin margins (Gennari et al. Reference Gennari, Lozar, Natalicchio, Zanella, Carnevale and Dela Pierre2020; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020). These sediments have recently been investigated to reconstruct the palaeoclimate and palaeohydrologic variability in the northern Mediterranean across the onset of the MSC (Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020). This previous study, which was based on inorganic geochemical proxies and on carbon and hydrogen stable isotope composition of lipids from terrestrial plant waxes, revealed fluctuations between more humid (shales) and more arid (marls) climates and an evolution towards moister conditions after the onset of the MSC (Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020).

Here, we focus on the same succession studied by Sabino et al. (Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020), integrating stratigraphic, petrographic, carbon and oxygen stable isotope analyses, and the study of molecular fossils to reconstruct the palaeoenvironments. This approach allows us to reconstruct the environmental changes across the onset of the MSC, and reveals how these changes influenced the stratigraphic architecture of the sediments deposited during times of change in the Mediterranean realm.

2. Geological setting

2.a. The Messinian succession in the Piedmont Basin

The Piedmont Basin (NW Italy; Fig. 1a) is a wedge-top basin located on the inner side of the SW Alpine arc filled with upper Eocene – Messinian sediments (Rossi et al. Reference Rossi, Mosca, Polino, Rogledi and Biffi2009; Mosca et al. Reference Mosca, Polino, Rogledi and Rossi2010). The Messinian succession is exposed on the uplifted southern and northern margins of the basin and starts with outer shelf to slope marls and shales (Sant’Agata Fossili Marls; Tortonian – lower Messinian; Sturani & Sampò, Reference Sturani and Sampò1973). These deposits, characterized by the repetition of shale and marl couplets, record progressively more restricted conditions heralding the onset of the MSC (Sturani, Reference Sturani and Drooger1973; Sturani & Sampò, Reference Sturani and Sampò1973). The lithological cyclicity was controlled by climate fluctuations driven by precession, with shales representing more humid conditions at precession minima and marls recording more arid conditions at precession maxima (Natalicchio et al. Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020). At the basin margins, this unit is overlain by shale and gypsum couplets belonging to the PLG unit (Fig. 1b). In settings of intermediate water depth, the gypsum beds of the lowermost PLG cycles are transitional to carbonate-rich layers and marls hosting fossilized microbial mats of putative sulphide-oxidizing bacteria, in turn passing in deep-water settings into dolomite-rich marls (Dela Pierre et al. Reference Dela Pierre, Clari, Bernardi, Natalicchio, Costa, Cavagna, Lozar, Lugli, Manzi, Roveri and Violanti2012, Reference Dela Pierre, Natalicchio, Lozar, Bonetto, Carnevale, Cavagna, Colombero, Sabino and Violanti2016; Natalicchio et al. Reference Natalicchio, Birgel, Peckmann, Lozar, Carnevale, Liu, Hinrichs and Dela Pierre2017, Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020; Fig. 1c). In the basin depocentre, only shales and marls are found (Irace et al. Reference Irace, Clemente, Natalicchio, Ossella, Trenkwalder, De Luca, Mosca, Piana, Polino and Violanti2009). The PLG unit and its deeper-water equivalents are overlain by chaotic and clastic gypsum facies (Valle Versa Chaotic complex; Irace et al. Reference Irace, Dela Pierre and Clari2005; Dela Pierre et al. Reference Dela Pierre, Festa and Irace2007), which are correlated to the Resedimented Lower Gypsum unit deposited during the second stage of the MSC (5.60–5.55 Ma; Roveri et al. Reference Roveri, Flecker, Krijgsman, Lofi, Lugli, Manzi, Sierro, Bertini, Camerlenghi, De Lange, Govers, Hilgen, Hübscher, Meijer and Stoica2014). The Messinian succession is completed by fluvio-deltaic and lacustrine sediments, referred to as the Cassano Spinola Conglomerates (Sturani, Reference Sturani1976; Dela Pierre et al. Reference Dela Pierre, Bernardi, Cavagna, Clari, Gennari, Irace, Lozar, Lugli, Manzi, Natalicchio, Roveri and Violanti2011, Reference Dela Pierre, Natalicchio, Lozar, Bonetto, Carnevale, Cavagna, Colombero, Sabino and Violanti2016; Fig. 1c), recording the third stage of the MSC (5.55–5.33 Ma; Roveri et al. Reference Roveri, Flecker, Krijgsman, Lofi, Lugli, Manzi, Sierro, Bertini, Camerlenghi, De Lange, Govers, Hilgen, Hübscher, Meijer and Stoica2014).

Fig. 1. (a) Distribution of the Messinian evaporites (pink) in the Western Mediterranean Basin and location of the Piedmont Basin (black box; modified from Manzi et al. Reference Manzi, Gennari, Hilgen, Krijgsman, Lugli, Roveri and Sierro2013). (b) Structural sketch map of the Piedmont Basin (modified from Bigi et al. Reference Bigi, Cosentino, Parotto, Sartori and Scandone1990); the star indicates the location of the Govone section. (c) Schematic profile of the Piedmont Basin, showing the stratigraphic architecture of the Messinian succession; the Govone section (vertical black bar) and the studied interval (red box) are indicated. Note that the gypsum beds are progressively younger towards the depocentre. MSC – Messinian salinity crisis; PB – Piedmont Basin; PLG – Primary Lower Gypsum (modified from Dela Pierre et al. Reference Dela Pierre, Bernardi, Cavagna, Clari, Gennari, Irace, Lozar, Lugli, Manzi, Natalicchio, Roveri and Violanti2011).

2.b. The Govone section and the position of the MSC onset

In the Govone section, located at the southern margin of the Piedmont Basin (44° 48′ 08″ N; 8° 07′ 34″ E; Fig. 1b), the whole Messinian succession is exposed, starting with 35 lithologic cycles up to 2 m in thickness (Gm1–Gm35; Bernardi et al. Reference Bernardi, Dela Pierre, Gennari, Lozar and Violanti2012; Bernardi, Reference Bernardi2013; Dela Pierre et al. Reference Dela Pierre, Natalicchio, Lozar, Bonetto, Carnevale, Cavagna, Colombero, Sabino and Violanti2016; Gennari et al. Reference Gennari, Lozar, Natalicchio, Zanella, Carnevale and Dela Pierre2020; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020; Fig. 2). These cycles consist of shale/marl couplets and belong to the Sant’Agata Fossili Marls. This unit is conformably overlain by the PLG unit (Fig. 1c), made up of nine cycles (Gg1–Gg9) of shales and gypsum-rich layers; the latter consist of flattened conical structures formed by millimetre-sized gypsum crystals enclosed in laminated gypsiferous silty mudstones (Bernardi, Reference Bernardi2013). The PLG is in turn overlain by shales and clastic evaporites belonging to the Valle Versa Chaotic complex, finally followed by fluvio-deltaic deposits of the Cassano Spinola Conglomerates (Fig. 1c; Bernardi, Reference Bernardi2013; Dela Pierre et al. Reference Dela Pierre, Natalicchio, Lozar, Bonetto, Carnevale, Cavagna, Colombero, Sabino and Violanti2016; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020).

Fig. 2. Tuning of the Govone section with the astronomical solution (65° N summer insolation; Laskar et al. Reference Laskar, Robutel, Joutel, Gastineau, Correia and Levrard2004) and correlation with the Perales section (Spain; Sierro et al. Reference Sierro, Hilgen, Krijgsman and Flores2001; Manzi et al. Reference Manzi, Gennari, Hilgen, Krijgsman, Lugli, Roveri and Sierro2013). Numbers in circles on the right represent the main bioevents reported in the main text. Bioz. – biozones; FAO – first abundant occurrence (modified from Gennari et al. Reference Gennari, Lozar, Natalicchio, Zanella, Carnevale and Dela Pierre2020; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020).

The age model for the Govone section (Gennari et al. Reference Gennari, Lozar, Natalicchio, Zanella, Carnevale and Dela Pierre2020) had already been adopted by Sabino et al. (Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020). The position of the MSC onset (5.97 Ma; Manzi et al. Reference Manzi, Gennari, Hilgen, Krijgsman, Lugli, Roveri and Sierro2013) was defined through the identification of diagnostic planktonic foraminifer bioevents, the ages of which were calibrated through the correlation with the astronomically tuned Perales section (Sierro et al. Reference Sierro, Hilgen, Krijgsman and Flores2001, Reference Sierro, Flores, Francés, Vázquez, Utrilla, Zamarreño, Erlenkeuser and Barcena2003; Manzi et al. Reference Manzi, Gennari, Hilgen, Krijgsman, Lugli, Roveri and Sierro2013; Fig. 2). The main bioevents identified by Gennari et al. (Reference Gennari, Lozar, Natalicchio, Zanella, Carnevale and Dela Pierre2020) are: (1) the first abundant occurrence (FAO) of Turborotalita multiloba in cycle Gm12, occurring in cycle UA15 in Perales and dated 6.415 Ma (Sierro et al. Reference Sierro, Hilgen, Krijgsman and Flores2001); (2) the left/right coiling change of Neogloboquadrina acostaensis in cycle Gm14, identified in Perales at cycle UA17 and dated 6.36 Ma (Sierro et al. Reference Sierro, Hilgen, Krijgsman and Flores2001); (3) the first influx of Globorotalia scitula in cycle Gm17, dated 6.29 Ma in the Perales section (cycle UA20; Sierro et al. Reference Sierro, Hilgen, Krijgsman and Flores2001); and (4) a second influx of Globorotalia scitula in cycle Gm24, which falls within an acme interval of T. multiloba. For the Perales section, Sierro et al. (Reference Sierro, Hilgen, Krijgsman and Flores2001) reported this acme between cycles UA23 and UA30, allowing Gennari et al. (Reference Gennari, Lozar, Natalicchio, Zanella, Carnevale and Dela Pierre2020) to correlate the second influx of G. scitula in the Govone section with the second influx recognized in cycle UA29 in the Perales section and dated at 6.10 Ma. Since the second influx of G. scitula occurs six precession cycles below the onset of the crisis in the Perales section (Sierro et al. Reference Sierro, Hilgen, Krijgsman and Flores2001), the MSC onset in Govone was placed at the base of the marls of cycle Gm30 (Gennari et al. Reference Gennari, Lozar, Natalicchio, Zanella, Carnevale and Dela Pierre2020; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020; Fig. 2), which accordingly correspond to the gypsum bed in cycle PLG1 of Perales (Fig. 2; Manzi et al. Reference Manzi, Gennari, Hilgen, Krijgsman, Lugli, Roveri and Sierro2013).

According to the adopted model, the uppermost six Sant’Agata Fossili Marls cycles below the first local gypsum bed (Gg1 cycle) represent the deep-water counterparts of the lowermost PLG cycles (Fig. 2), with marls representing the time equivalents of the shallow-water marginal gypsum (Gennari et al. Reference Gennari, Lozar, Natalicchio, Zanella, Carnevale and Dela Pierre2020; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020). In this study we investigate four pre-MSC (Gm26–Gm29) and four MSC (Gm30–Gm33) cycles, representing a time interval of c. 150 ka. Younger strata of the Govone section containing gypsum (Fig. 1c) are not studied here.

3. Materials and methods

3.a. Petrography and mineralogy

A total of 33 fresh, unweathered samples were excavated from cycles Gm26 to Gm33 (on average, 4 samples per cycle); the Govone sedimentary rocks tend to be well preserved and continuously exposed due to ongoing erosion in the river bed. A total of 18 slabs and thin-sections were obtained from representative samples (10 from pre-MSC cycles and 8 from MSC cycles) and studied at the Department of Earth Sciences, University of Turin by optical (transmitted, reflected and UV light) and scanning electron microscopy (SEM). SEM analyses were performed on carbon-coated stubs for morphological analyses and on carbon-coated, polished thin-sections for semi-quantitative elemental analysis and backscattered electron imagery, using a JEOL JSM IT300LV scanning electron microscope equipped with an energy-dispersive EDS Oxford Instrument Link System microprobe (University of Turin).

X-ray diffraction (XRD) analyses were performed on all 33 samples using a Panalytical X’Pert PRO diffractometer (CuKα radiation, 40 kV, 40 mA, step size 0.0167, 5 s per step) at the Department of Geodynamics and Sedimentology, University of Vienna. The samples were loaded in the sample holders as oriented powders. The X-ray diffraction patterns were interpreted using the Panalytical software ‘X’Pert High score plus’ to determine the carbonate phase mineralogy.

3.b. Carbon and oxygen stable isotopes

Carbon (δ13C) and oxygen (δ18O) bulk-rock stable isotope analyses were performed on 9 samples (pre-MSC cycles, 1 sample; MSC cycles, 8 samples). These analyses add to those of Bernardi (Reference Bernardi2013) who analysed 20 samples, 12 from pre-MSC cycles and 8 from MSC cycles. Analyses were performed at the MARUM stable isotope laboratory (University of Bremen). The 9 new samples were measured at 75°C on a Finnigan MAT 252 gas isotope ratio mass spectrometer connected to a Kiel III automated carbonate preparation device. The instrument was calibrated against an in-house standard (ground Solnhofen limestone), which in turn was calibrated against the NBS 19 calcite standard. Over the measurement period, the standard deviations of the in-house standard were 0.04‰ for δ13C and 0.06‰ for δ18O values. Data are reported in delta-notation versus V-PDB. When dolomite was the only carbonate phase, the δ18O values were corrected for −0.8‰ for measurements from Bernardi (Reference Bernardi2013; analyses performed at 50°C; Sharma & Clayton, Reference Sharma and Clayton1965) and −1.2‰ for the new measurements (analyses performed at 75°C; Rosenbaum & Sheppard, Reference Rosenbaum and Sheppard1986; Kim et al. Reference Kim, Mucci and Taylor2007). The correction was necessary to account for the fractionation effect during the phosphoric acid reaction (see online Supplementary Material, available at http://journals.cambridge.org/geo, for further details).

3.c. Total inorganic and organic carbon contents

Total inorganic (TIC) and organic (TOC) carbon contents were measured at the Institute for Geology of the University of Hamburg. After drying (at 50°C for 24 h), the samples were manually ground with an agate mortar. The powders were split in two aliquots: one fraction was heated to 1350°C and total carbon (TC) contents were measured using a LECO SC-144DR Carbon Analyser equipped with an infrared detector. The second fraction was first heated to 550°C for 5 h to remove organic carbon (OC) and then heated to 1350°C to measure TIC contents. Prior to and after sample analyses, a Synthetic Carbon Leco 502-029 (1.01 ± 0.02 carbon%) standard was measured. TOC contents were determined using the formula TOC = TC–TIC.

3.d. Lipid biomarker analyses

Lipid biomarker analyses were performed on 21 samples (at least 2 samples per cycle) using the procedure described in Sabino et al. (Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020). Briefly, after a modified Bligh and Dyer extraction, samples were separated in an n-hexane-soluble and a dichloromethane-soluble fraction. The former was further separated into four sub-fractions: (a) hydrocarbons, (b) ketones, (c) alcohols and (d) carboxylic acids. The alcohol fraction was derivatized for 1 h at 70°C by adding pyridine and N,O-bis(trimethylsilyl)trifluoroacetamide (BSTFA) (1:1; v/v). The ketones and carboxylic acids did not contain any indigenous compounds and are not discussed further. Alcohol and hydrocarbon fractions were dried and re-dissolved in n-hexane for analyses. Compounds were identified using a Thermo Scientific Trace gas chromatograph (GC) Ultra coupled to a Thermo Scientific DSQ II mass spectrometer (MS) through comparison of retention times and published mass spectra. Quantification was achieved with a Fisons Instruments GC 8000 series equipped with a flame-ionization detector (FID). Internal standards used for quantification were 5α-cholestane for the hydrocarbon fraction and 1-nonadecanol and DAGE C18-18 for the alcohol fraction. The carrier gases were helium and hydrogen for the GC-MS and GC-FID analyses, respectively. Both devices were equipped with an Agilent HP-5MS UI fused silica column with a length of 30 m, a diameter of 0.25 mm and a film thickness of 0.25 μm. The GC temperature programme was: 50°C (3 min); from 50°C to 230°C (held 2 min) at 25°C/min; then from 230°C to 320°C (held 20 min) at 6°C/min.

An aliquot of the hydrocarbon fraction was used to isolate branched and cyclic compounds from the n-alkanes. The aliquot was treated with 300 mg of 5 Å molecular sieve and 2 mL cyclohexane, was shaken well, then extracted by ultrasonication for 2 h. The extract was filtered and the mole sieve was repeatedly washed with cyclohexane and filtered once again to release all branched and cyclic compounds. After drying with molecular nitrogen, the extract was re-dissolved in n-hexane for compound identification and quantification. The GC temperature program was: 50°C (3 min); from 50°C to 230°C (held 2 min) at 25°C/min; then from 230°C to 325°C (held 25 min) at 6°C/min.

3.d.1. Desulphurization with nickel boride

Desulphurization was performed for the dichloromethane-soluble fractions (asphaltenes), applying a procedure slightly modified from Schouten et al. (Reference Schouten, Pavlović, Sinninghe Damsté and de Leeuw1993) and Blumenberg et al. (Reference Blumenberg, Mollenhauer, Zabel, Reimer and Thiel2010). Briefly, the asphaltenes were dissolved in 8 mL tetrahydrofuran/methanol (1:1, v/v), then 200 mg of each anhydrous nickel chloride and sodium borohydride were added. After 1 h reaction time, the samples were centrifuged and the supernatant was collected. The solid residue was extracted twice with dichloromethane/methanol (1:1, v/v), then centrifuged again and the supernatant was combined with the previous extract. An aqueous solution was added and the organic layer collected and dried with a rotary evaporator. The residual organic phase was filtered through dry sodium sulphate to completely remove water. Column chromatography was performed using a silica gel column and 2.5 mL of n-hexane/dichloromethane (9:1, v/v) as eluent to separate the released hydrocarbons after desulphurization from the polar compounds. The desulphurized hydrocarbon fraction was re-dissolved in n-hexane and the compounds identified through GC-MS analyses using the following program: from 60°C (1 min) to 150°C at 15°C/min and then up to 320°C (held 40 min) at 4°C/min. Due to low contents, the desulphurized hydrocarbon fractions were also run in single ion mode (SIM) with the masses m/z 133 and 546, specific masses of di-aromatic carotenoids and isorenieratane, respectively.

3.d.2. Glycerol dibiphytanyl glycerol tetraether analyses

Glycerol dibiphytanyl glycerol tetraethers (GDGTs) were obtained from 10% vol. of the total lipid extract (TLE) according to the method of Hopmans et al. (Reference Hopmans, Schouten, Pancost, van der Meer and Sinninghe Damsté2000) and modified by Baumann et al. (Reference Baumann, Taubner, Bauersachs, Steiner, Schleper, Peckmann, Rittmann and Birgel2018). The aliquot was dissolved in n-hexane and an internal standard (C46 GDGT; 12 mg/L) was added. The analyses were performed using a Varian MS Workstation 6.91 high-performance liquid chromatography (HPLC) system coupled to a Varian 1200 L triple quadrupole mass spectrometer. The compounds were separated on a Grace Prevail Cyano column (150 × 2.1 mm; 3 μm particle size) and a guard column, held at 30°C. The following gradient was applied: linear change from 97.5% A (100% n-hexane) and 2.5% B (90% n-hexane: 10% 2-propanol; v/v) to 75% A and 25% B from 0 to 35 min; then linearly to 100% B in 5 min and held for 8 min; and thereafter, back to 97.5% A and 2.5% B to re-equilibrate the column for 12 min. The total run time was 60 min and the solvent flow was kept constant at 0.3 mL/min during the entire run time. The identification of GDGTs was achieved using a mass spectrometer equipped with an atmospheric pressure chemical ionization (APCI) interface operated in positive ion mode. The APCI parameters were: molecular nitrogen as nebulizing gas with a pressure of 60 psi; temperature fluctuating between 35°C and 40°C; 50°C as API housing temperature; 200°C for the drying gas of the API with a pressure of 12 psi; and 400°C and 18 psi for the APCI auxiliary gas temperature and pressure, respectively. The injection volume was 10 µL. The scanned spectral range was set at m/z 500 to 750 and 950 to 1500. The response factors were evaluated after every 4 to 5 samples using a standard mixture containing synthetic archaeol (1,2-Di-O-phytanyl-sn-glycerol; CAS 99341-19-2), DAGE C18:18 (CAS 6076-38-6), DAGE C18:18-4ene (1,3-Dilinoleoyl-rac-glycerol; CAS 15818-46-9) and synthetic C46 GDGT (CAS 138456-87-8). The response factors between synthetic archaeol and C46 GDGT were usually around 1.5:1. Relative lipid abundances were determined by selecting individual base peaks with the target mass m/z, including ions symmetrically with ±1.0 of target m/z.

4. Results

4.a. Petrography and mineralogy

The studied interval is characterized by 8 lithological cycles of shales and argillaceous marls (referred to as marls) up to 1 m thick (Fig. 3a, b). The inorganic carbon contents range from 0.4% to 6.0%, with higher contents in marls (Fig. 3a). XRD analyses indicate that the inorganic carbon in the Govone section is composed of calcite and dolomite (see online Supplementary Material, available at http://journals.cambridge.org/geo). Dolomite coexists with calcite in cycles Gm26 and Gm27, while it represents the only carbonate phase in cycle Gm28 (except for sample Gm28.1) and above (Fig. 3a).

Fig. 3. (a) The Govone section with sample locations, total inorganic carbon contents, carbonate mineralogy and bulk-rock carbon (δ13C) and oxygen (δ18O) stable isotope values. White squares and circles refer to δ13C and δ18O values from Bernardi (Reference Bernardi2013), whereas black squares and circles refer to values from this study. (b) Outcrop view of the Govone section. The red line corresponds to the onset of the MSC; the white dashed lines indicate the top of marl beds; insets indicate the position of figures (c) and (d). (c) Close-up of pre-MSC laminated organic-rich shales. (d) Close-up of pre-MSC homogenous marls.

4.a.1. Shales

Shales are laminated, dark-green-coloured (Fig. 3c) and rich in organic matter (TOC > 2%). They are characterized by an alternation of discontinuous, sub-millimetric, cream-coloured (type A) and dark-grey-coloured (type B) laminae, irregularly interrupted by layers up to 1.5 cm thick composed of terrigenous silt- to sand-sized grains (Fig. 4a). These layers are considered to represent the products of more intense fluvial discharge to the basin (Dela Pierre et al. Reference Dela Pierre, Clari, Natalicchio, Ferrando, Giustetto, Lozar, Lugli, Manzi, Roveri and Violanti2014).

Fig. 4. Organic-rich shales. (a) Polished slab cut perpendicular to bedding. Type A and B laminae separated by coarser terrigenous layers (t.l.). The black dashed lines envelope a type A lamina, the white arrows indicate two type B laminae. (b) Fluorescent, lens-shaped faecal pellets (black oval) and flattened, faintly fluorescent aggregate (white arrow); UV-fluorescence image. (c) Irregular aggregates cut parallel to bedding; UV-fluorescence image. (d) Faecal pellet (dashed line) composed of coccoliths; SEM micrograph. (e) Irregular aggregate (dashed line) mostly composed of dolomite microcrystals from the MSC part of the section. The white box indicates the position of (f); SEM micrograph. (f) Globular dolomite microcrystals and scattered pyrite framboids (white arrows); SEM micrograph. (g) Hollow dolomite microcrystals (black arrow and inset) and scattered pyrite framboids (white arrows); SEM micrograph. (h) Planktonic foraminifer test (black arrow) surrounded by coccoliths (white arrows). The inset shows a detail with coccoliths and a small pyrite framboid (arrow); SEM micrograph. (i) Foraminifer test partially filled with pyrite (p); plane-polarized light. The inset is a backscattered electron image of a pyrite infilling, consisting of an aggregate of tiny pyrite framboids. Scale bar in the inset is 20 μm.

Type A laminae show a distinct bulbous to pinch-and-swell-type character (Fig. 4a; sensu Pilskaln & Pike, Reference Pilskaln and Pike2001). They are formed of two varieties of fluorescent peloids: (1) flattened to lens-shaped peloids up to 0.5 mm across with well-defined boundaries; and (2) irregular aggregates, up to 1 mm across, with irregular, diffused boundaries (Fig. 4a–c). In cycles Gm26 and Gm27, the peloids are composed of scattered silt-sized siliciclastic grains (quartz, mica flakes and feldspars), black organic matter particles and coccoliths, very abundant in the lens-shaped peloids (Fig. 4d). In cycles Gm28 to Gm33, peloids are barren of coccoliths and mostly formed of dolomite microcrystals up to 10 µm across, scattered pyrite framboids and accessory silt-sized siliciclastic grains (Fig. 4e). Dolomite microcrystals show prevalently globular, hemispherical or cauliflower-like (sensu Warthmann et al. Reference Warthmann, van Lith, Vasconcelos, McKenzie and Karpoff2000) habits (Fig. 4e–g), and sometimes exhibit a central hollow (Fig. 4g).

Type B laminae are only a few microns thick, laterally discontinuous (Fig. 4a) and composed of silt- to clay-sized terrigenous grains (mainly quartz, feldspars and mica flakes), black amorphous organic matter debris enriched in sulphur (as revealed by EDS analyses) and abundant pyrite framboids, up to 10 µm across. These laminae are interpreted to represent the product of enhanced terrigenous input into the basin, under the control of short-term, probably seasonal, climate change (Dela Pierre et al. Reference Dela Pierre, Clari, Natalicchio, Ferrando, Giustetto, Lozar, Lugli, Manzi, Roveri and Violanti2014).

Both types of laminae contain scattered planktonic foraminifer tests (only in cycles Gm26 and Gm27; Fig. 4h, i), fish scales and vertebrae. Tiny (<10 μm) pyrite framboids (Fig. 4f–h), locally filling foraminifer tests (Fig. 4i), are abundant.

4.a.2. Marls

Marls are grey-coloured (Fig. 3d) with bioturbation only present in the pre-MSC cycles (Fig. 5a). Marls are rich in silt-sized terrigenous grains (mainly quartz, feldspar and mica flakes), and contain finely dispersed pyrite grains up to 10 µm across (framboids and octahedral crystals; Fig. 5b, c). Dolomite is common in the MSC cycles, consisting of globular and cauliflower-like (Fig. 5d–f) microcrystals, up to 20 µm across, often typified by a central hollow (Fig. 5e, f). Dumbbell-shaped crystals also occur (Fig. 5f).

Fig. 5. (a) Photomicrographs and (b–f) SEM images of marls. (a) Bioturbation traces in pre-MSC marl (black box); the inset is a close-up image in UV light of the bioturbation traces, highlighted by dashed line. Pyrite: (b) framboid, (c) octahedral crystals. Dolomite microcrystals with (d) cauliflower, (e) globular and (f, arrow) dumbbell-like shapes. The arrows in (e) indicate central hollows in dolomite microcrystals.

4.b. Carbon and oxygen stable isotopes

The bulk-rock δ13C values range from −5.2‰ to +0.3‰, lacking an obvious relationship with lithology (Fig. 3; online Supplementary Material, available at http://journals.cambridge.org/geo). However, a general trend towards more negative δ13C values is observed across the MSC onset, evidenced by overall more pronounced 13C-depletion in MSC cycles (average δ13C = −2.4‰) than in pre-MSC cycles (average δ13C = −1.0‰). A moderate negative correlation was found when plotting the bulk-rock δ13C values of shales (n = 6) and marls (n = 7) against the TIC contents, where dolomite was the only carbonate phase detected (shales, r = −0.7; marls, r = −0.6; online Supplementary Material). The δ18O bulk-rock values vary from −4.5‰ to +5.5‰. As for δ13C values, a correlation with lithology was not observed (Fig. 3). The largest fluctuations were found in the pre-MSC cycles (+5.4‰ ≥ δ18O ≥ −4.5‰), while the MSC cycles show less variable, positive δ18O values (average +3.5‰; Fig. 3).

4.c. Total organic carbon contents and lipid biomarkers

The TOC contents vary from 1.0% to 3.1% (Fig. 6; online Supplementary Material, available at http://journals.cambridge.org/geo) and follow lithological cyclicity, with higher contents in shales (average 2.4%) and lower contents in marls (average 1.6%), particularly in pre-MSC cycles (as low as 1.0%).

Fig. 6. TOC contents, lipid biomarker distributions and caldarchaeol/crenarchaeol ratio (cald/cren ratio) across the onset of the Messinian salinity crisis in the Govone section. Note pentacyclic C30 sulphide (Pentac. sulphide) and isorenieratane (isorenier.) in the tetrahymanol distribution profile and the non-metric scale in the lycopene distribution profile. The red line at the base of marls in cycle Gm30 indicates the onset of the Messinian salinity crisis. GDGT – glycerol dibiphytanyl glycerol tetraethers; Ly – lycopane; THM – tetrahymanol; GDGT-0 – caldarchaeol; GDGT-5 – crenarchaeol.

4.c.1. Archaeol and GDGT distribution and caldarchaeol/crenarchaeol ratio

GDGTs and C20-20 archaeol (archaeol) were detected in the Govone samples. The former are mainly represented by GDGT-0 (caldarchaeol) and GDGT-5 (crenarchaeol), with relative contents ranging from 18.0% to 31.2% and 18.3% to 32.9% of the total GDGTs plus archaeol assemblage, respectively (Fig. 6, Table 1; online Supplementary Material, available at http://journals.cambridge.org/geo). GDGTs with 1–3 cyclopentane rings and GDGT-5’ (crenarchaeol isomer) yielded relative contents of ≤11.0% on average (Fig. 6, Table 1). Archaeol revealed relative contents between 7.8% and 42.9% of the total GDGTs plus archaeol inventory (Fig. 6, Table 1, online Supplementary Material). The distribution of GDGTs and archaeol did not vary across the MSC onset (Fig. 6, Table 1). The caldarchaeol/crenarchaeol ratio varies between 0.8 and 1.5, with an average of 1.1 (Fig. 6; online Supplementary Material).

Table 1. Archaeol, GDGTs (both in % relative to all GDGTs plus archaeol), lycopane and tetrahymanol abundances in pre-MSC and MSC Govone sediments. Max = maximum content, Min = minimum content, GDGT-0 = caldarchaeol, GDGT-5 = crenarchaeol.

4.c.2. Acyclic and cyclic triterpenoids and the lycopane/n-C31 alkane ratio

The hydrocarbon and alcohol fractions contain, among other compounds (n-alkanes and n-alkanols, steranes and sterols, hopanes and hopanols, long-chain diols and keto-ols; thiolanes, thianes and thiophenes; data not shown), the acyclic triterpenoid lycopane and the pentacyclic triterpenoid tetrahymanol.

Lycopane, isolated with a molecular sieve from the n-alkanes, was found in most of the 18 samples (except for marls in cycle Gm26), yielding varying contents (up to 25.0 µg/g TOC). In general, shales revealed much higher contents than marls (Fig. 6, Table 1; online Supplementary Material, available at http://journals.cambridge.org/geo). Lycopane contents increase upwards, especially above the MSC onset (Fig. 6). The lycopane/n-C31 alkane ratio, a palaeo-oxicity proxy introduced by Sinninghe Damsté et al. (Reference Sinninghe Damsté, Kuypers, Schouten, Schulte and Rullkötter2003), ranges from 0.01 to 1.2 (online Supplementary Material) with most ratios < 0.3; such values suggest that the seafloor was constantly oxygenated (cf. Sinninghe Damsté et al. Reference Sinninghe Damsté, Kuypers, Schouten, Schulte and Rullkötter2003), although petrographic and other organic geochemical evidence suggests the opposite (see Section 5.c below). This inconsistency most likely results from the high flux of leaf-wax-derived long-chain n-alkanes relative to lycopane, which can significantly affect the ratio and make the use of its absolute values unreliable (cf. Sinninghe Damsté et al. Reference Sinninghe Damsté, Kuypers, Schouten, Schulte and Rullkötter2003). We therefore decided not to apply the absolute values of the ratio, as suggested by Sinninghe Damsté et al. (Reference Sinninghe Damsté, Kuypers, Schouten, Schulte and Rullkötter2003), but refer to trends between the lithologies to infer relative variations in palaeo-oxicity.

Tetrahymanol occurs in all samples with contents ranging from 1.1 to 26.5 µg/g TOC (Fig. 6; online Supplementary Material, available at http://journals.cambridge.org/geo). Tetrahymanol is more abundant in shales than in marls, although the difference between the two lithologies is less pronounced in the MSC interval than in the pre-MSC part of the section (Fig. 6, Table 1; online Supplementary Material).

4.c.3. Hydrocarbons liberated after desulphurization

The desulphurization of asphaltenes released mostly n-alkanes (C14 to C40), a pentacyclic C30 sulphide (cf. Poinsot et al. Reference Poinsot, Schneckenburger, Adam, Schaeffer, Trendel, Riva and Albrecht1998) and, especially in shales, phytane. As well as these compounds, we identified C27 to C29 steranes, C31 to C35 hopanes, and, only in samples from cycles Gm28 to Gm30, the diaromatic carotenoid isorenieratane. In most cases (with the exception of sample Gm30), isorenieratane was identified in single ion mode (SIM) only (m/z 133 and 546). This is in accordance with very low contents, rendering quantification impossible.

5. Discussion

5.a. Water-column stratification and euxinia across the onset of the MSC

The restriction of the Mediterranean leading to the MSC was associated with an intensification of water-column stratification (Roveri et al. Reference Roveri, Flecker, Krijgsman, Lofi, Lugli, Manzi, Sierro, Bertini, Camerlenghi, De Lange, Govers, Hilgen, Hübscher, Meijer and Stoica2014) that persisted during the entire MSC event (Christeleit et al. Reference Christeleit, Brandon and Zhuang2015; Simon & Meijer, Reference Simon and Meijer2017; García-Veigas et al. Reference García-Veigas, Cendón, Gibert, Lowenstein and Artiaga2018). The same evolution has been reconstructed for the Piedmont Basin (Dela Pierre et al. Reference Dela Pierre, Bernardi, Cavagna, Clari, Gennari, Irace, Lozar, Lugli, Manzi, Natalicchio, Roveri and Violanti2011; Bernardi, Reference Bernardi2013; Violanti et al. Reference Violanti, Lozar, Natalicchio, Dela Pierre, Bernardi, Clari and Cavagna2013; Lozar et al. Reference Lozar, Violanti, Bernardi, Dela Pierre and Natalicchio2018), especially by the study of the Pollenzo section, a section about 20 km SW from the Govone section (Natalicchio et al. Reference Natalicchio, Birgel, Peckmann, Lozar, Carnevale, Liu, Hinrichs and Dela Pierre2017, Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019). For the Pollenzo section, the most compelling evidence of increasing water-column stratification after the MSC onset is the appearance of tetrahymanol. This compound is produced by organisms commonly thriving at chemoclines (Wakeham et al. Reference Wakeham, Turich, Schubotz, Podlaska, Li, Varela, Astor, Sáenz, Rush, Sinninghe Damsté, Summons, Scranton, Taylor and Hinrichs2012), namely bacterivorous ciliates, anoxygenic phototrophic bacteria and aerobic methanotrophic bacteria (Kleemann et al. Reference Kleemann, Poralla, Englert, Kjøsen, Liaaen-Jensen, Neunlist and Rohmer1990; Harvey & McManus, Reference Harvey and McManus1991; Rashby et al. Reference Rashby, Sessions, Summons and Newman2007; Eickhoff et al. Reference Eickhoff, Birgel, Talbot, Peckmann and Kappler2013; Banta et al. Reference Banta, Wei and Welander2015; Cordova-Gonzales et al. 2020); tetrahymanol is therefore considered a robust indicator of water-column stratification (e.g. Reference Sinninghe Damsté, Kenig, Koopmans, Köster, Schouten, Hayes and de LeeuwSinninghe Damsté et al. 1995b ).

Tetrahymanol was found throughout the section at Govone, suggesting that, in the more distal settings of the Piedmont Basin, stratified conditions had already been established before the onset of the MSC. The higher tetrahymanol contents in shales suggest that stratification was more intense at precession minima, most likely in response of enhanced riverine runoff (cf. Natalicchio et al. Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020). Enhanced riverine runoff in turn promoted episodes of high productivity and phytoplankton blooms. The latter are testified by the very abundant lens-shaped peloids and irregular aggregates, which have been interpreted as faecal pellets and marine snow-flakes, respectively (cf. Dela Pierre et al. Reference Dela Pierre, Clari, Natalicchio, Ferrando, Giustetto, Lozar, Lugli, Manzi, Roveri and Violanti2014), both main components of so-called marine snow (e.g. Alldredge & Silver, Reference Alldredge and Silver1988; Alldredge et al. Reference Alldredge, Cowles, MacIntyre, Rines, Donaghay, Greenlaw, Holliday, Dekshenieks, Sullivan and Zaneveld2002). Interestingly, in the shales of the uppermost pre-MSC cycles (Gm28-30), tetrahymanol is accompanied by isorenieratane, derived from the degradation of isorenieratene sourced by green sulphur bacteria (Repeta et al. Reference Repeta, Simpson, Jorgensen and Jannasch1989; Van Gemerden & Mas, Reference Van Gemerden, Mas, Blankenship, Madigan and Bauer1995). These phototrophic bacteria require hydrogen sulphide to perform anoxygenic photosynthesis (e.g. Van Gemerden & Mas, Reference Van Gemerden, Mas, Blankenship, Madigan and Bauer1995) and are found in modern stratified basins when sulphidic conditions extend into the photic zone (e.g. Black Sea, Sinninghe Damsté et al. Reference Sinninghe Damsté, Wakeham, Kohnen, Hayes and de Leeuw1993; Wakeham et al. Reference Wakeham, Amann, Freeman, Hopmans, Jørgensen, Putnam, Schouten, Sinninghe Damsté, Talbot and Woebken2007; Cariaco Basin, Wakeham et al. Reference Wakeham, Turich, Schubotz, Podlaska, Li, Varela, Astor, Sáenz, Rush, Sinninghe Damsté, Summons, Scranton, Taylor and Hinrichs2012), making isorenieratene and its derivatives biomarkers for photic zone euxinia (e.g. Repeta & Simpson, Reference Repeta and Simpson1991; Kuypers et al. Reference Kuypers, Pancost, Nijenhuis and Sinninghe Damsté2002). The co-occurrence of tetrahymanol and isorenieratane in shales therefore supports pronounced water-column stratification at precession minima already before the onset of the MSC, occasionally accompained by photic zone euxinia. The lower contents of tetrahymanol and the lack of isorenieratane in the MSC shales apparently suggest the weakening of stratification, which is however in contrast with evidence from the neighbouring Pollenzo section (Natalicchio et al. Reference Natalicchio, Birgel, Peckmann, Lozar, Carnevale, Liu, Hinrichs and Dela Pierre2017, Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019). This inconsistency could reflect unfavourable environmental conditions at precession minima for the producers of isorenieratene and tetrahymanol after the MSC onset in the distal part of the basin (Govone), rather than the weakening of water-column stratification. Interestingly, in modern stratified basins, ciliates are found to feed, among other organisms, on green sulphur bacteria (Wakeham et al. Reference Wakeham, Amann, Freeman, Hopmans, Jørgensen, Putnam, Schouten, Sinninghe Damsté, Talbot and Woebken2007, Reference Wakeham, Turich, Schubotz, Podlaska, Li, Varela, Astor, Sáenz, Rush, Sinninghe Damsté, Summons, Scranton, Taylor and Hinrichs2012). The trophic relationship between autotrophic green sulphur bacteria and heterotrophic ciliates, already inferred for the MSC sediments of the Northern Apennines (Vena del Gesso Basin, Sinninghe Damsté et al. Reference Sinninghe Damsté, Frewin, Kenig and de Leeuw1995a ), can potentially explain the trend of compound contents in the Govone section, where tetrahymanol peaks coincide with the presence of isorenieratane. Adverse conditions for green sulphur bacteria were most likely related to an intensified input of terrigenous material by rivers after the MSC onset, as evidenced by increasing contents of terrestrial plant waxes (Natalicchio et al. Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020). Enhanced riverine runoff might have increased turbidity and input of oxygen in the upper water column, both deleterious factors for phototrophic green sulphur bacteria (Van Gemerden & Mas, Reference Van Gemerden, Mas, Blankenship, Madigan and Bauer1995), in turn causing a decrease of the ciliate population.

Evidence of persistent stratification and sulphidic conditions in the water column at precession minima after the MSC onset, although not extending into the photic zone, is provided by the occurrence of small pyrite framboids and sulphur-enriched organic matter. The small size of the super-abundant pyrite framboids (< 10 μm) may suggest that they formed within an euxinic water column (cf. Wilkin et al. Reference Wilkin, Barnes and Brantley1996; Passier et al. Reference Passier, Middelburg, de Lange and Böttcher1997; Bond & Wignall, Reference Bond and Wignall2010; Tagliavento et al. Reference Tagliavento, Lauridsen and Stemmerik2020). Organic sulphur compounds, such as the pentacyclic C30 sulphide, are believed to form during a very early diagenetic stage, possibly already in the water column when reduced-sulphur species (e.g. hydrogen sulphide) exceed reduced iron and consequently bind to settling organic matter (Sinninghe Damsté & de Leeuw, Reference Sinninghe Damsté and de Leeuw1990; Wakeham et al. Reference Wakeham, Sinninghe Damsté, Kohnen and de Leeuw1995; Poinsot et al. Reference Poinsot, Schneckenburger, Adam, Schaeffer, Trendel, Riva and Albrecht1998).

Compared with shales, marls reveal lower contents of tetrahymanol. Such a pattern reflects the weakening of stratification at precession maxima in response to more efficient mixing of the water column (e.g. Sierro et al. Reference Sierro, Hilgen, Krijgsman and Flores2001; Violanti et al. Reference Violanti, Lozar, Natalicchio, Dela Pierre, Bernardi, Clari and Cavagna2013). Mixing was promoted by drier climate conditions and reduced fluvial discharge, in accordance with higher element/Al ratios and a decrease of contents in the sediments of terrestrial plant waxes, which show deuterium enrichment (Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020). However, water-column stratification never completely ceased to exist, since the occurrence of small pyrite framboids and sulphur organic compounds suggests that sulphidic conditions were intermittently present in the water column. The trend to slightly higher tetrahymanol contents in MSC marls compared with pre-MSC marls suggests that water-column stratification was more intense at precession maxima after the onset of the MSC compared with precession maxima during the pre-MSC, in agreement with the appearance of tetrahymanol in time-equivalent more marginal sediments (Pollenzo section; Natalicchio et al. Reference Natalicchio, Birgel, Peckmann, Lozar, Carnevale, Liu, Hinrichs and Dela Pierre2017). This was possibly due to the establishment of a wetter climate in the northern Mediterranean (cf. Natalicchio et al. Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020).

5.b. Persistence of marine conditions in surface waters after the onset of the MSC

The Govone section is characterized by the progressive decline and final disappearance of calcareous plankton approaching the MSC onset (Bernardi, Reference Bernardi2013). Such ecological change is recorded in other sections across the Mediterranean and was taken as evidence of progressively harsher conditions in surface waters, with fluctuations of salinity up to high levels not tolerated by most marine biota (e.g. Sierro et al. Reference Sierro, Flores, Zamarreño, Vázquez, Utrilla, Francés, Hilgen and Krijgsman1999; Blanc-Valleron et al. Reference Blanc-Valleron, Pierre, Caulet, Caruso, Rouchy, Cespuglio, Sprovieri, Pestrea and Di Stefano2002; Manzi et al. Reference Manzi, Roveri, Gennari, Bertini, Biffi, Giunta, Iaccarino, Lanci, Lugli, Negri, Riva, Rossi and Taviani2007). The establishment of hypersaline conditions during the MSC is also reflected by the archaeal di- and tetraether lipid assemblages of MSC carbonates from Sicily and Calabria (the so-called ‘Calcare di Base’); in these deposits, C20-20 archaeol, C20-25 archaeol (extended archaeol) and caldarchaeol dominate the isoprenoid diphytanyl glycerol diether (DGD) and GDGT inventories (Turich & Freeman, Reference Turich and Freeman2011; Birgel et al. Reference Birgel, Guido, Liu, Hinrichs, Gier and Peckmann2014).

In contrast, the isoprenoid DGD and GDGT distribution found for the Govone section is significantly different from that of the Calcare di Base and does not agree with widespread hypersalinity after the onset of the MSC. In fact, archaeol does not dominate over the total GDGTs (c. 25% of the total GDGT and DGD lipid inventory) and is never accompanied by extended archaeol, allowing the exclusion of a significant contribution from halophilic archaea (cf. Teixidor et al. Reference Teixidor, Grimait, Pueyo and Rodriguez-Valera1993; Dawson et al. Reference Dawson, Freeman and Macalady2012). Possible alternative sources for archaeol are non-halophilic archaea, such as methanogens, methanotrophs and marine Euryarchaeota thriving in epi- to mesopelagic zones (De Rosa & Gambacorta, Reference De Rosa and Gambacorta1988; Blumenberg et al. Reference Blumenberg, Seifert, Reitner, Pape and Michaelis2004; Turich et al. Reference Turich, Freeman, Bruns, Conte, Jones and Wakeham2007; Sollai et al. Reference Sollai, Villanueva, Hopmans, Keil and Sinninghe Damsté2019). Methanogenic and methanotrophic archaea, on the other hand, can be excluded for the Govone section, since the distribution of isoprenoidal di- and tetraether lipids varies from those observed in methane-seep environments, the latter typified by the dominance of archaeol, hydroxyarchaeol and GDGT-1 and 2 over the other dibiphytanyl tetraethers (De Rosa & Gambacorta, Reference De Rosa and Gambacorta1988; Blumenberg et al. Reference Blumenberg, Seifert, Reitner, Pape and Michaelis2004; Wakeham et al. Reference Wakeham, Amann, Freeman, Hopmans, Jørgensen, Putnam, Schouten, Sinninghe Damsté, Talbot and Woebken2007). Consequently, marine Euryarchaeota thriving in the epi- to mesopelagic zones of a normal marine water column are the most likely source of archaeol (cf. Turich et al. Reference Turich, Freeman, Bruns, Conte, Jones and Wakeham2007; Schouten et al. Reference Schouten, Hopmans, Baas, Boumann, Standfest, Könneke, Stahl and Sinninghe Damsté2008; Elling et al. Reference Elling, Könneke, Mußmann, Greve and Hinrichs2015; Sollai et al. Reference Sollai, Villanueva, Hopmans, Keil and Sinninghe Damsté2019). Interestingly, it has been suggested that mesopelagic marine Euryarchaeota from the group III (MGIII) are important producers of archaeol, especially in oxygen-deficient zones where they can even dominate the archaeal community (Belmar et al. Reference Belmar, Molina and Ulloa2011; Sollai et al. Reference Sollai, Villanueva, Hopmans, Keil and Sinninghe Damsté2019). Since episodical oxygen deficiency seems to have characterized the water column of the Piedmont Basin during the MSC, mesopelagic marine Euryarchaeota are the most likely source of archaeol in the Govone section, indicating the persistence of marine rather than hypersaline conditions across the onset of the MSC.

Marine conditions are in accordance with the distribution of the GDGTs and the caldarchaeol/crenarchaeol ratio (c. 1), which mirror the distribution and the ratios reported for open-ocean settings (e.g. Turich et al. Reference Turich, Freeman, Bruns, Conte, Jones and Wakeham2007; Kim et al. Reference Kim, van der Meer, Schouten, Helmke, Willmott, Sangiorgi, Koç, Hopmans and Sinninghe Damsté2010; Pearson & Ingalls, Reference Pearson and Ingalls2013; Schouten et al. Reference Schouten, Hopmans and Sinninghe Damsté2013). Under such conditions, the main contributors to the GDGT pool are planktonic marine Euryarchaeota of the group II (MGII) and Thaumarchaeota (marine group I, MGI), dwelling in the epi- to mesopelagic zones (Schouten et al. Reference Schouten, Hopmans and Sinninghe Damsté2013; Santoro et al. Reference Santoro, Richter and Dupont2019). Despite some controversy (Lincoln et al. Reference Lincoln, Wai, Eppley, Church, Summons and DeLong2014; Schouten et al. Reference Schouten, Villanueva, Hopmans, van der Meer and Sinninghe Damsté2014), MGII archaea are thought to source mostly acyclic GDGTs (caldarchaeol), whereas MGI archaea synthesize both acyclic and polycyclic GDGTs, especially crenarchaeol (Elling et al. Reference Elling, Könneke, Nicol, Stieglmeier, Bayer, Spieck, de la Torre, Becker, Thomm, Prosser, Herndl, Schleper and Hinrichs2017; Zeng et al. Reference Zeng, Liu, Farley, Wei, Metcalf, Summons and Welander2019). This pattern results in the dominance of caldarchaeol and crenarchaeol over the other GDGTs and a caldarchaeol/crenarchaeol ratio of c. 1 (Sinninghe Damsté et al. Reference Sinninghe Damsté, Schouten, Hopmans, van Duin and Geenevasen2002; Schouten et al. Reference Schouten, Hopmans and Sinninghe Damsté2013 for a review), which closely mirror the tetraether archaeal lipid distribution and the ratio of marine planktonic Thaumarchaeota growing in cultures at normal salinity (c. 35‰ salinity; Elling et al. Reference Elling, Könneke, Mußmann, Greve and Hinrichs2015).

The persistence of marine conditions in surface waters challenges the idea of an establishment of a completely hypersaline water mass early on during the MSC (e.g. Bellanca et al. Reference Bellanca, Caruso, Ferruzza, Neri, Rouchy, Sprovieri and Blanc-Valleron2001). The lack of pervasive hypersaline conditions agrees with (1) the presence of marine fossils (foraminifers, calcareous nannoplankton, diatoms, fishes) above the MSC onset in more marginal sections of the Piedmont Basin (Violanti et al. Reference Violanti, Lozar, Natalicchio, Dela Pierre, Bernardi, Clari and Cavagna2013; Dela Pierre et al. Reference Dela Pierre, Clari, Natalicchio, Ferrando, Giustetto, Lozar, Lugli, Manzi, Roveri and Violanti2014; Lozar et al. Reference Lozar, Violanti, Bernardi, Dela Pierre and Natalicchio2018; Carnevale et al. Reference Carnevale, Gennari, Lozar, Natalicchio, Pellegrino and Dela Pierre2019) and (2) palaeosalinity estimates from gypsum fluid inclusions, indicating that the parent brine had a salinity possibly even lower than seawater (Natalicchio et al. Reference Natalicchio, Dela Pierre, Lugli, Lowenstein, Feiner, Ferrando, Manzi, Roveri and Clari2014). In this light, the disappearance of calcareous plankton in the Govone section two cycles below the onset of the MSC most likely reflects diagenetic dissolution of calcareous skeletons rather than harsh environmental conditions in superficial waters (Dela Pierre et al. Reference Dela Pierre, Clari, Natalicchio, Ferrando, Giustetto, Lozar, Lugli, Manzi, Roveri and Violanti2014).

5.c. Fluctuations in oxygen concentration of bottom waters across the onset of the MSC

Intensification of water-column stratification in the Mediterranean Basin shortly before the onset of the MSC was associated with widespread depletion of oxygen in bottom waters, which is recorded by the precession-driven deposition of organic-rich sediments (Hilgen & Krijgsman, Reference Hilgen and Krijgsman1999; Krijgsman et al. Reference Krijgsman, Fortuin, Hilgen and Sierro2001, Reference Krijgsman, Blanc-Valleron, Flecker, Hilgen, Kouwenhoven, Merle, Orszag-Sperber and Rouchy2002) and by the replacement of oxyphilic benthic foraminifera with stress-tolerant taxa (e.g. Kouwenhoven et al. Reference Kouwenhoven, Seidenkrantz and van der Zwaan1999; Blanc-Valleron et al. Reference Blanc-Valleron, Pierre, Caulet, Caruso, Rouchy, Cespuglio, Sprovieri, Pestrea and Di Stefano2002; Gennari et al. Reference Gennari, Lozar, Turco, Dela Pierre, Lugli, Manzi, Natalicchio, Roveri, Schreiber and Taviani2018). Such a trend was confirmed for the Piedmont Basin by changes in the assemblages of benthic foraminifers in the pre-MSC sediments (Bernardi, Reference Bernardi2013; Violanti et al. Reference Violanti, Lozar, Natalicchio, Dela Pierre, Bernardi, Clari and Cavagna2013). However, very little information is available for the microfossil-poor or microfossil-barren MSC strata.

The Govone section archives the fluctuation of oxygen contents at the seafloor driven by precessional forcing. Humid climate and enhanced riverine runoff at precession minima (Natalicchio et al. Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020) favoured intense water-column stratification and bottom anoxia, which is witnessed by the deposition of laminated, organic-rich shales (TOC > 2%). Evidence of bottom-water anoxia during the deposition of shales includes their high contents of lycopane. Studies of modern (Wakeham et al. Reference Wakeham, Freeman, Pease and Hayes1993; Sinninghe Damsté et al. Reference Sinninghe Damsté, Kuypers, Schouten, Schulte and Rullkötter2003) and ancient sediments (e.g. Freeman et al. Reference Freeman, Hayes, Trendel and Albrecht1990; Kenig et al. Reference Kenig, Sinninghe Damsté, Frewin, Hayes and De Leeuw1995; Behrooz et al. Reference Behrooz, Naafs, Dickson, Love, Batenburg and Pancost2018; Löhr et al. Reference Löhr, Kennedy, George, Williamson and Xu2018) have shown that preservation of lycopane is highest when bottom waters are anoxic.

Lower TOC contents (mostly < 2%) and the sharp drop of lycopane contents (≤0.4 µg/g TOC) in marls indicate higher oxygen levels at the seafloor than at times of shale deposition, reflecting precession-driven climate change (Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020). However, changes can be observed across the onset of the MSC, possibly recording different degrees of bottom-water ventilation before and after the advent of the crisis. In particular, pre-MSC marls show bioturbation traces and contain few stress-tolerant benthic foraminifers (Bernardi, Reference Bernardi2013; Dela Pierre et al. Reference Dela Pierre, Natalicchio, Lozar, Bonetto, Carnevale, Cavagna, Colombero, Sabino and Violanti2016), suggesting that the seafloor was at least episodically oxygenated. In contrast, the absence of bioturbation and the slight increase of lycopane contents in MSC marls compared with pre-MSC marls indicate that bottom waters became progressively oxygen depleted, agreeing with an intensification of water-column stratification at precession maxima during the MSC relative to precession maxima in pre-MSC times (Section 5.a above). Despite such a trend to lower oxygen levels, the only low lycopane contents at precession maxima compared with precession minima (shale deposition) suggest that bottom waters were not fully anoxic during the deposition of marls (cf. Sinninghe Damsté et al. Reference Sinninghe Damsté, Kuypers, Schouten, Schulte and Rullkötter2003).

5.d. Primary Lower Gypsum: implications on depositional environments and stratigraphic architecture

The integration of geochemical, sedimentological and petrographic data indicates that during the early phase of the MSC the water column in the Piedmont Basin was stratified and comprised (1) an oxygenated upper layer, typified by mostly marine conditions and receiving input of freshwater from rivers (Natalicchio et al. Reference Natalicchio, Dela Pierre, Birgel, Brumsack, Carnevale, Gennari, Gier, Lozar, Pellegrino, Sabino, Schnetger and Peckmann2019; Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020) and/or low-salinity water from the Paratethys (Grothe et al. Reference Grothe, Andreetto, Reichart, Wolthers, Van Baak, Vasiliev, Stoica, Sangiorgi, Middelburg, Davies and Krijgsman2020); and (2) an anoxic to euxinic lower layer, formed by denser, more saline waters (Dela Pierre et al. Reference Dela Pierre, Clari, Natalicchio, Ferrando, Giustetto, Lozar, Lugli, Manzi, Roveri and Violanti2014; Natalicchio et al. Reference Natalicchio, Birgel, Peckmann, Lozar, Carnevale, Liu, Hinrichs and Dela Pierre2017), the latter agreeing with more positive δ18O values (average +3.5‰) of MSC sediments. The two water masses were separated by a pycnocline, at which chemical gradients established with time and a chemocline developed, as observed for modern basins (Wakeham et al. Reference Wakeham, Amann, Freeman, Hopmans, Jørgensen, Putnam, Schouten, Sinninghe Damsté, Talbot and Woebken2007, Reference Wakeham, Turich, Schubotz, Podlaska, Li, Varela, Astor, Sáenz, Rush, Sinninghe Damsté, Summons, Scranton, Taylor and Hinrichs2012) and reconstructed for ancient basins with stratified water masses (e.g. the Badenian basin of Eastern Europe; Babel, Reference Babel2004; Babel & Bogucki, Reference Babel, Bogucki, Schreiber, Lugli and Babel2007).

Stratigraphic data reveal a lateral facies change in the PLG unit from the margin to the depocentre of the Piedmont Basin for sediments deposited during precession maxima (Dela Pierre et al. Reference Dela Pierre, Bernardi, Cavagna, Clari, Gennari, Irace, Lozar, Lugli, Manzi, Natalicchio, Roveri and Violanti2011). Bottom-grown selenite gypsum is observed only in the shallow, marginal part of the basin (Arnulfi section; Fig. 7). Moving towards the depocentre, gypsum makes lateral transition to marls and carbonates with filamentous fossils (Pollenzo section; Fig. 7) – interpreted to represent chemotrophic microbial mats (Dela Pierre et al. Reference Dela Pierre, Clari, Bernardi, Natalicchio, Costa, Cavagna, Lozar, Lugli, Manzi, Roveri and Violanti2012) – and finally to dolomite-rich marls (Govone section; Fig. 7). Conversely, at precession minima, organic-rich shales were deposited across the entire Piedmont Basin (Dela Pierre et al. Reference Dela Pierre, Bernardi, Cavagna, Clari, Gennari, Irace, Lozar, Lugli, Manzi, Natalicchio, Roveri and Violanti2011, Reference Dela Pierre, Natalicchio, Lozar, Bonetto, Carnevale, Cavagna, Colombero, Sabino and Violanti2016; Fig. 7). We suggest that these stratigraphic patterns reflect vertical chemocline oscillations controlled by precession (Fig. 7).

Fig. 7. Reconstruction of the water-column structure of the Piedmont Basin at (a) precession minima, insolation maxima and shale deposition and (b) precession maxima, insolation minima and marl deposition during the earliest phase of the Messinian salinity crisis. The black arrows indicate the positions of sections (Ar: Arnulfi; Pz: Pollenzo; Gv: Govone). The thickness of the chemocline is emphasized in (b) to highlight the different conditions described in the text.

The widespread seafloor anoxia recorded by shales indicates that the chemocline was located above the sea bottom across most of the basin during precession minima. These conditions favoured organic matter preservation in the water column and led to an increased sedimentation of organic matter, which, in turn, favoured heterotrophic, bacterial sulphate reduction and promoted bacterially mediated precipitation of early diagenetic dolomite in anoxic, organic-rich sediments (Fig. 7a). This inference agrees with (1) the negative bulk-rock δ13C values (cf. Petrash et al. Reference Petrash, Bialik, Bontognali, Vasconcelos, Roberts, McKenzie and Konhauser2017); (2) the moderate negative correlation (r = −0.7) between TIC contents, represented by dolomite, and bulk-rock δ13C values (high TIC contents coincide with low δ13C values); and (3) dolomite crystal habits (Figs 4f, g, 5d–f; cf. Warthmann et al. Reference Warthmann, van Lith, Vasconcelos, McKenzie and Karpoff2000; van Lith et al. Reference van Lith, Warthmann, Vasconcelos and McKenzie2003; Sanz-Montero et al. Reference Sanz-Montero, Rodríguez-Aranda and García Del Cura2008; Bontognali et al. Reference Bontognali, Vasconcelos, Warthmann, Bernasconi, Dupraz, Strohmenger and McKenzie2010; Han et al. Reference Han, Schultz, Zhang, Zhu, Meng and Geesey2016).

The change towards a less humid climate at precession maxima (Sabino et al. Reference Sabino, Schefuß, Natalicchio, Dela Pierre, Birgel, Bortels, Schnetger and Peckmann2020) resulted in the deepening of the chemocline and the thinning of the anoxic zone (Fig. 7b). Under these circumstances, selenite grew only in the shallower part of the basin (Arnulfi section, Fig. 7b), where low-salinity waters enriched in calcium and sulphate ions, possibly derived from leaching of former evaporites (Natalicchio et al. Reference Natalicchio, Dela Pierre, Lugli, Lowenstein, Feiner, Ferrando, Manzi, Roveri and Clari2014), occurred above an oxygenated sea bottom (e.g. García-Veigas et al. Reference García-Veigas, Cendón, Gibert, Lowenstein and Artiaga2018). In more distal settings, the absence of gypsum was possibly related to low-oxygen conditions, which favoured bacterial sulphate reduction and resultant gypsum undersaturation (Fig. 7b; cf. Babel, Reference Babel2004; de Lange & Krijgsman, Reference de Lange and Krijgsman2010; García-Veigas et al. Reference García-Veigas, Cendón, Gibert, Lowenstein and Artiaga2018). The parts of the basin where the chemocline impinged on the seafloor (Pollenzo section, Fig. 7b; Natalicchio et al. Reference Natalicchio, Birgel, Peckmann, Lozar, Carnevale, Liu, Hinrichs and Dela Pierre2017) were covered by microbial mats of putative sulphide-oxidizing bacteria (Natalicchio et al. Reference Natalicchio, Birgel, Peckmann, Lozar, Carnevale, Liu, Hinrichs and Dela Pierre2017). In modern settings, these bacteria thrive under hypoxic conditions where the chemocline intercepts the seafloor and steep gradients between electron acceptors and hydrogen sulphide occur (e.g. the Black Sea; Jessen et al. Reference Jessen, Lichtschlag, Struck and Boetius2016). In contrast, in deeper parts of the basin (Govone section; Fig. 7b) the sea bottom was in contact with more reducing waters. These conditions favoured the preservation of organic matter in the water column, enhanced the deposition of organic matter at the seafloor and led to an increase in bacterial sulphate reduction in the sediments, which triggered the widespread precipitation of early diagenetic microbial dolomite (Fig. 7b).

6. Conclusions

Hydrological change affecting the Piedmont Basin during the early stages of the Messinian salinity crisis (MSC) led to an intensification of water-column stratification. The water column was divided into denser, more saline and oxygen-depleted bottom waters separated by a chemocline from an upper water layer consisting of oxic seawater influenced by freshwater inflow. No evidence of a sharp increase of salinity across the MSC onset was found. Vertical oscillations of the chemocline exerted control over the stratigraphic architecture of the Primary Lower Gypsum unit and its deeper-water equivalents deposited during the first stage of the MSC. This study documents how temporal and spatial changes of water masses with different redox chemistries must be carefully considered when interpreting the MSC event.

Acknowledgements

We thank S. Beckmann (Universität Hamburg) for technical support during organic geochemical analyses, L. Baumann (Universität Hamburg) for support in the application of the GDGTs proxies, D. Bortels (Universität Hamburg) for help with lipid biomarker extraction, J. Richarz (Universität Hamburg) for the analysis of TOC contents and H. Kuhnert (University of Bremen) for the bulk-rock carbon and oxygen stable isotopes analyses and help with data processing. S. Cavagna (University of Torino) is thanked for support with SEM-EDS analyses. This project received funding from University of Torino grants (ex 60% 2017 and 2019) to F. Dela Pierre. M. Sabino is funded by a doctoral scholarship provided by the Landesgraduiertenförderung of the state of Hamburg. The article is based upon work towards COST (European Cooperation in Science and Technology) Action ‘Uncovering the Mediterranean salt giant’ (MEDSALT). Comments by Iuliana Vasiliev and an anonymous reviewer helped to improve the manuscript.

Declaration of Interest

None.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/S0016756820000874

References

Alldredge, AL and Silver, MW (1988) Characteristics, dynamics and significance of marine snow. Progress in Oceanography 20, 4182.CrossRefGoogle Scholar
Alldredge, AL, Cowles, TJ, MacIntyre, S, Rines, JEB, Donaghay, PL, Greenlaw, CF, Holliday, DV, Dekshenieks, MM, Sullivan, JM and Zaneveld, JRV (2002) Occurrence and mechanisms of formation of a dramatic thin layer of marine snow in a shallow Pacific fjord. Marine Ecology Progress Series 233, 112.CrossRefGoogle Scholar
Babel, M (2004) Models for evaporite, selenite and gypsum microbialite deposition in ancient saline basins. Acta Geologica Polonica 54, 219–49.Google Scholar
Babel, M and Bogucki, A (2007) The Badenian evaporite basin of the northern Carpathian Foredeep as a model of a meromictic selenite basin. In Evaporites Through Space and Time (eds Schreiber, BC, Lugli, S and Babel, M, pp. 219–46. Geological Society of London, Special Publication no. 285.CrossRefGoogle Scholar
Banta, AB, Wei, JH and Welander, PV (2015) A distinct pathway for tetrahymanol synthesis in bacteria. Proceedings of the National Academy of Sciences 112, 13478–83.CrossRefGoogle Scholar
Baumann, LMF, Taubner, RS, Bauersachs, T, Steiner, M, Schleper, C, Peckmann, J, Rittmann, SK-MR and Birgel, D (2018) Intact polar lipid and core lipid inventory of the hydrothermal vent methanogens Methanocaldococcus villosus and Methanothermococcus okinawensis . Organic Geochemistry 126, 3342.CrossRefGoogle Scholar
Behrooz, L, Naafs, BDA, Dickson, AJ, Love, GD, Batenburg, SJ and Pancost, RD (2018) Astronomically driven variations in depositional environments in the South Atlantic during the Early Cretaceous. Paleoceanography and Paleoclimatology 33, 894912.CrossRefGoogle Scholar
Bellanca, A, Caruso, A, Ferruzza, G, Neri, R, Rouchy, JM, Sprovieri, M and Blanc-Valleron, MM (2001) Transition from marine to hypersaline conditions in the Messinian Tripoli Formation from the marginal areas of the central Sicilian Basin. Sedimentary Geology 140, 87105.CrossRefGoogle Scholar
Belmar, L, Molina, V and Ulloa, O (2011) Abundance and phylogenetic identity of archaeoplankton in the permanent oxygen minimum zone of the eastern tropical South Pacific. FEMS Microbiology Ecology 78, 314–26.CrossRefGoogle ScholarPubMed
Bernardi, E (2013) Integrated stratigraphy of the northernmost record of the Messinian salinity crisis: new insights from the Tertiary Piedmont Basin. PhD thesis, University of Torino, Torino, Italy. Published thesis.Google Scholar
Bernardi, E, Dela Pierre, F, Gennari, R, Lozar, F and Violanti, D (2012) Astrochronological calibration and paleoenvironmental reconstruction of the Messinian events at the Northern edge of the Mediterranean: The Govone section (Tertiary Piedmont Basin). Rendiconti Online Società Geologica Italiana 20, 1011.Google Scholar
Bigi, G, Cosentino, D, Parotto, M, Sartori, R and Scandone, P (1990) Structural model of Italy: Geodinamic Project. Consiglio Nazionale delle Ricerche (S.EL.CA, scale 1:500,000, sheet 1).Google Scholar
Birgel, D, Guido, A, Liu, X, Hinrichs, K-U, Gier, S and Peckmann, J (2014) Hypersaline conditions during deposition of the Calcare di Base revealed from archaeal di- and tetraether inventories. Organic Geochemistry 77, 1121.CrossRefGoogle Scholar
Blanc-Valleron, M-M, Pierre, C, Caulet, JP, Caruso, A, Rouchy, J-M, Cespuglio, G, Sprovieri, R, Pestrea, S and Di Stefano, E (2002) Sedimentary, stable isotope and micropaleontological records of paleoceanographic change in the Messinian Tripoli Formation (Sicily, Italy). Palaeogeography, Palaeoclimatology, Palaeoecology 185, 255–86.CrossRefGoogle Scholar
Blumenberg, M, Mollenhauer, G, Zabel, M, Reimer, A and Thiel, V (2010) Decoupling of bio- and geohopanoids in sediments of the Benguela Upwelling System (BUS). Organic Geochemistry 41, 1119–29.CrossRefGoogle Scholar
Blumenberg, M, Seifert, R, Reitner, J, Pape, T and Michaelis, W (2004) Membrane lipid patterns typify distinct anaerobic methanotrophic consortia. Proceedings of the National Academy of Sciences 101, 11111–16.CrossRefGoogle ScholarPubMed
Bond, DPG and Wignall, PB (2010) Pyrite framboid study of marine Permian-Triassic boundary sections: a complex anoxic event and its relationship to contemporaneous mass extinction. Bulletin of the Geological Society of America 122, 1265–79.CrossRefGoogle Scholar
Bontognali, TRR, Vasconcelos, C, Warthmann, RJ, Bernasconi, SM, Dupraz, C, Strohmenger, CJ and McKenzie, JA (2010) Dolomite formation within microbial mats in the coastal sabkha of Abu Dhabi (United Arab Emirates). Sedimentology 57, 824–44.CrossRefGoogle Scholar
Camerlenghi, A and Aloisi, V (2020) Uncovering the Mediterranean Salt Giant (MEDSALT) - Scientific networking as incubator of cross-disciplinary research in Earth Sciences. European Review 28, 4061.CrossRefGoogle Scholar
Capella, W, Flecker, R, Hernández-Molina, FJ, Simon, D, Meijer, PT, Rogerson, M, Sierro, FJ and Krijgsman, W (2019) Mediterranean isolation preconditioning the Earth System for late Miocene climate cooling. Scientific Reports 9, article 3795.CrossRefGoogle ScholarPubMed
Carnevale, G, Gennari, R, Lozar, F, Natalicchio, M, Pellegrino, L and Dela Pierre, F (2019) Living in a deep desiccated Mediterranean Sea: an overview of the Italian fossil record of the Messinian salinity crisis. Bollettino della Società Paleontologica Italiana 58, 109–40.Google Scholar
Christeleit, EC, Brandon, MT and Zhuang, G (2015) Evidence for deep-water deposition of abyssal Mediterranean evaporites during the Messinian salinity crisis. Earth and Planetary Science Letters 427, 226–35.CrossRefGoogle Scholar
Cordova-Gonzalez, A, Birgel, D, Kappler, A and Peckmann, J (2020) Carbon stable isotope patterns of cyclic terpenoids: a comparison of cultured alkaliphilic aerobic methanotrophic bacteria and methane-seep environments. Organic Geochemistry 139, article 103940.CrossRefGoogle Scholar
Dawson, KS, Freeman, KH and Macalady, JL (2012) Molecular characterization of core lipids from halophilic archaea grown under different salinity conditions. Organic Geochemistry 48, 18.CrossRefGoogle Scholar
de Lange, GJ and Krijgsman, W (2010) Messinian salinity crisis: a novel unifying shallow gypsum/deep dolomite formation mechanism. Marine Geology 275, 273–7.CrossRefGoogle Scholar
De Rosa, M and Gambacorta, A (1988) The lipids of archaebacteria. Progress in Lipid Research 27, 153–75.CrossRefGoogle ScholarPubMed
Dela Pierre, F, Bernardi, E, Cavagna, S, Clari, P, Gennari, R, Irace, A, Lozar, F, Lugli, S, Manzi, V, Natalicchio, M, Roveri, M and Violanti, D (2011) The record of the Messinian salinity crisis in the Tertiary Piedmont Basin (NW Italy): the Alba section revisited. Palaeogeography, Palaeoclimatology, Palaeoecology 310, 238–55.CrossRefGoogle Scholar
Dela Pierre, F, Clari, P, Bernardi, E, Natalicchio, M, Costa, E, Cavagna, S, Lozar, F, Lugli, S, Manzi, V, Roveri, M and Violanti, D (2012) Messinian carbonate-rich beds of the Tertiary Piedmont Basin (NW Italy): microbially-mediated products straddling the onset of the salinity crisis. Palaeogeography, Palaeoclimatology, Palaeoecology 344–345, 7893.CrossRefGoogle Scholar
Dela Pierre, F, Clari, P, Natalicchio, M, Ferrando, S, Giustetto, R, Lozar, F, Lugli, S, Manzi, V, Roveri, M and Violanti, D (2014) Flocculent layers and bacterial mats in the mudstone interbeds of the Primary Lower Gypsum unit (Tertiary Piedmont basin, NW Italy): archives of palaeoenvironmental changes during the Messinian salinity crisis. Marine Geology 355, 7187.CrossRefGoogle Scholar
Dela Pierre, F, Festa, A and Irace, A (2007) Interaction of tectonic, sedimentary, and diapiric processes in the origin of chaotic sediments: an example from the Messinian of Torino Hill (Tertiary Piedmont Basin, northwestern Italy). Bulletin of the Geological Society of America 119, 1107–19.CrossRefGoogle Scholar
Dela Pierre, F, Natalicchio, M, Lozar, F, Bonetto, S, Carnevale, G, Cavagna, S, Colombero, S, Sabino, M and Violanti, D (2016) The northernmost record of the Messinian salinity crisis (Piedmont basin, Italy). Geological Field Trips 8 (2.1), 158.CrossRefGoogle Scholar
Eickhoff, M, Birgel, D, Talbot, HM, Peckmann, J and Kappler, A (2013) Oxidation of Fe(II) leads to increased C-2 methylation of pentacyclic triterpenoids in the anoxygenic phototrophic bacterium Rhodopseudomonas palustris strain TIE-1. Geobiology 11, 268–78.CrossRefGoogle ScholarPubMed
Elling, FJ, Könneke, M, Mußmann, M, Greve, A and Hinrichs, K-U (2015) Influence of temperature, pH, and salinity on membrane lipid composition and TEX86 of marine planktonic thaumarchaeal isolates. Geochimica et Cosmochimica Acta 171, 238–55.CrossRefGoogle Scholar
Elling, FJ, Könneke, M, Nicol, GW, Stieglmeier, M, Bayer, B, Spieck, E, de la Torre, JR, Becker, KW, Thomm, M, Prosser, J I, Herndl, GJ, Schleper, C and Hinrichs, K-U (2017) Chemotaxonomic characterisation of the thaumarchaeal lipidome. Environmental Microbiology 19, 2681–700.CrossRefGoogle ScholarPubMed
Flecker, R, Krijgsman, W, Capella, W, de Castro Martíns, C, Dmitrieva, E, Mayser, JP, Marzocchi, A, Modestu, S, Ochoa, D, Simon, D, Tulbure, M, van den Berg, B, van der Schee, M, de Lange, G, Ellam, R, Govers, R, Gutjahr, M, Hilgen, F, Kouwenhoven, T, Lofi, J, Meijer, P, Sierro, FJ, Bachiri, N, Barhoun, N, Alami, AC, Chacon, B, Flores, JA, Gregory, J, Howard, J, Lunt, D, Ochoa, M, Pancost, R, Vincent, S and Yousfi, MZ (2015) Evolution of the Late Miocene Mediterranean-Atlantic gateways and their impact on regional and global environmental change. Earth-Science Reviews 150, 365–92.CrossRefGoogle Scholar
Freeman, KH, Hayes, JM, Trendel, JM and Albrecht, P (1990) Evidence from carbon isotope measurements for diverse origins of sedimentary hydrocarbons. Nature 343, 254–6.CrossRefGoogle ScholarPubMed
García-Veigas, J, Cendón, D I, Gibert, L, Lowenstein, TK and Artiaga, D (2018) Geochemical indicators in Western Mediterranean Messinian evaporites: implications for the salinity crisis. Marine Geology 403, 197214.CrossRefGoogle Scholar
Gennari, R, Lozar, F, Natalicchio, M, Zanella, E, Carnevale, G, and Dela Pierre, F (2020). Chronology of the Messinian events in the northernmost part of the Mediterranean: the Govone section (Piedmont Basin, NW Italy). Rivista Italiana di Paleontologia e Stratigrafia 126, 517–60.Google Scholar
Gennari, R, Lozar, F, Turco, E, Dela Pierre, F, Lugli, S, Manzi, V, Natalicchio, M, Roveri, M, Schreiber, BC and Taviani, M (2018) Integrated stratigraphy and paleoceanographic evolution of the pre-evaporitic phase of the Messinian salinity crisis in the Eastern Mediterranean as recorded in the Tokhni section (Cyprus island). Newsletters on Stratigraphy 51, 3355.CrossRefGoogle Scholar
Grothe, A, Andreetto, F, Reichart, G-J, Wolthers, M, Van Baak, CGC, Vasiliev, I, Stoica, M, Sangiorgi, F, Middelburg, JJ, Davies, GR and Krijgsman, W (2020) Paratethys pacing of the Messinian Salinity Crisis: low salinity waters contributing to gypsum precipitation? Earth and Planetary Science Letters 532, article 116029.CrossRefGoogle Scholar
Han, X, Schultz, L, Zhang, W, Zhu, J, Meng, F and Geesey, GG (2016) Mineral formation during bacterial sulfate reduction in the presence of different electron donors and carbon sources. Chemical Geology 435, 4959.CrossRefGoogle Scholar
Harvey, HR and McManus, GB (1991) Marine ciliates as a widespread source of tetrahymanol and hopan-3β-ol in sediments. Geochimica et Cosmochimica Acta 55, 3387–90.CrossRefGoogle Scholar
Hilgen, FJ and Krijgsman, W (1999) Cyclostratigraphy and astrochronology of the Tripoli diatomite formation (pre-evaporite Messinian, Sicily, Italy). Terra Nova 11, 1622.Google Scholar
Hopmans, EC, Schouten, S, Pancost, RD, van der Meer, MTJ and Sinninghe Damsté, JS (2000) Analysis of intact tetraether lipids in archaeal cell material and sediments by high performance liquid chromatography/atmospheric pressure chemical ionization mass spectrometry. Rapid Communications in Mass Spectrometry 14, 585–9.3.0.CO;2-N>CrossRefGoogle ScholarPubMed
Hsü, KJ, Ryan, WBF and Cita, MB (1973) Late Miocene desiccation of the Mediterranean. Nature 242, 240–4.CrossRefGoogle Scholar
Irace, A, Clemente, P, Natalicchio, M, Ossella, L, Trenkwalder, S, De Luca, DA, Mosca, P, Piana, F, Polino, R and Violanti, D (2009) Geologia e idrostratigrafia profonda della Pianura Padana Occidentale. Firenze: La Nuova Lito, 110 p.Google Scholar
Irace, A, Dela Pierre, F and Clari, P (2005) «Normal» and «chaotic» deposits in the Messinian Gessoso-solfifera Fm. at the north-eastern border of the Langhe domain (Tertiary Piedmont basin). Bollettino della Società Geologica Italiana 4, 7785.Google Scholar
Isaji, Y, Kawahata, H, Takano, Y, Ogawa, NO, Kuroda, J, Yoshimura, T, Lugli, S, Manzi, V, Roveri, M and Ohkouchi, N (2019a) Diazotrophy drives primary production in the organic-rich shales deposited under a stratified environment during the Messinian salinity crisis (Vena Del Gesso, Italy). Frontiers in Earth Science 7, article 85.CrossRefGoogle Scholar
Isaji, Y, Yoshimura, T, Kuroda, J, Tamenori, Y, Jiménez-Espejo, FJ, Lugli, S, Manzi, V, Roveri, M, Kawahata, H and Ohkouchi, N (2019b) Biomarker records and mineral compositions of the Messinian halite and K–Mg salts from Sicily. Progress in Earth and Planetary Science 6, article 60.CrossRefGoogle Scholar
Jessen, GL, Lichtschlag, A, Struck, U and Boetius, A (2016) Distribution and composition of thiotrophic mats in the hypoxic zone of the Black Sea (150-170 m water depth, Crimea margin). Frontiers in Microbiology 7, article 1011.CrossRefGoogle Scholar
Kenig, F, Sinninghe Damsté, JS, Frewin, NL, Hayes, JM and De Leeuw, JW (1995) Molecular indicators for palaeoenvironmental change in a Messinian evaporitic sequence (Vena del Gesso, Italy). II: high-resolution variations in abundances and 13C contents of free and sulphur-bound carbon skeletons in a single marl bed. Organic Geochemistry 23, 485526.CrossRefGoogle Scholar
Kim, J-H, van der Meer, J, Schouten, S, Helmke, P, Willmott, V, Sangiorgi, F, Koç, N, Hopmans, EC and Sinninghe Damsté, JS (2010) New indices and calibrations derived from the distribution of crenarchaeal isoprenoid tetraether lipids: implications for past sea surface temperature reconstructions. Geochimica et Cosmochimica Acta 74, 4639–54.CrossRefGoogle Scholar
Kim, ST, Mucci, A and Taylor, BE (2007) Phosphoric acid fractionation factors for calcite and aragonite between 25 and 75°C: revisited. Chemical Geology 246, 135–46.CrossRefGoogle Scholar
Kleemann, G, Poralla, K, Englert, G, Kjøsen, H, Liaaen-Jensen, S, Neunlist, S and Rohmer, M (1990) Tetrahymanol from the phototrophic bacterium Rhodopseudomonas palustris: first report of a gammacerane triterpene from a prokaryote. Journal of General Microbiology 136, 2551–3.CrossRefGoogle Scholar
Kouwenhoven, TJ, Seidenkrantz, M-S and van der Zwaan, GJ (1999) Deep-water changes: the near-synchronous disappearance of a group of benthic foraminifera from the Late Miocene Mediterranean. Palaeogeography, Palaeoclimatology, Palaeoecology 152, 259–81.CrossRefGoogle Scholar
Krijgsman, W, Blanc-Valleron, M-M, Flecker, R, Hilgen, FJ, Kouwenhoven, TJ, Merle, D, Orszag-Sperber, F and Rouchy, J-M (2002) The onset of the Messinian salinity crisis in the Eastern Mediterranean (Pissouri Basin, Cyprus). Earth and Planetary Science Letters 194, 299310.CrossRefGoogle Scholar
Krijgsman, W, Capella, W, Simon, D, Hilgen, FJ, Kouwenhoven, TJ, Meijer, PT, Sierro, FJ, Tulbure, MA, van den Berg, BCJ, van der Schee, M and Flecker, R (2018) The Gibraltar Corridor: Watergate of the Messinian Salinity Crisis. Marine Geology 403, 238–46.CrossRefGoogle Scholar
Krijgsman, W, Fortuin, AR, Hilgen, FJ and Sierro, FJ (2001) Astrochronology for the Messinian Sorbas basin (SE Spain) and orbital (precessional) forcing for evaporite cyclicity. Sedimentary Geology 140, 4360.CrossRefGoogle Scholar
Kuypers, MMM, Pancost, RD, Nijenhuis, IA and Sinninghe Damsté, JS (2002) Enhanced productivity led to increased organic carbon burial in the euxinic North Atlantic basin during the late Cenomanian oceanic anoxic event. Paleoceanography 17, article 1051.CrossRefGoogle Scholar
Laskar, J, Robutel, P, Joutel, F, Gastineau, M, Correia, ACM and Levrard, B (2004) A long-term numerical solution for the insolation quantities of the Earth. Astronomy & Astrophysics 428, 261–85.CrossRefGoogle Scholar
Lincoln, SA, Wai, B, Eppley, JM, Church, MJ, Summons, RE and DeLong, EF (2014) Planktonic Euryarchaeota are a significant source of archaeal tetraether lipids in the ocean. Proceedings of the National Academy of Sciences 111, 9858–63.CrossRefGoogle ScholarPubMed
Löhr, SC, Kennedy, MJ, George, SC, Williamson, RJ and Xu, H (2018) Sediment microfabric records mass sedimentation of colonial cyanobacteria and extensive syndepositional metazoan reworking in Pliocene sapropels. The Depositional Record 4, 293317.CrossRefGoogle Scholar
Lozar, F, Violanti, D, Bernardi, E, Dela Pierre, F and Natalicchio, M (2018) Identifying the onset of the Messinian salinity crisis: a reassessment of the biochronostratigraphic tools (Piedmont Basin, NW Italy). Newsletters on Stratigraphy 51, 1131.CrossRefGoogle Scholar
Manzi, V, Gennari, R, Hilgen, F, Krijgsman, W, Lugli, S, Roveri, M and Sierro, FJ (2013) Age refinement of the Messinian salinity crisis onset in the Mediterranean. Terra Nova 25, 315–22.CrossRefGoogle Scholar
Manzi, V, Lugli, S, Roveri, M, Schreiber, BC and Gennari, R (2011) The Messinian ‘Calcare di Base’ (Sicily, Italy) revisited. Bulletin of the Geological Society of America 123, 347–70.CrossRefGoogle Scholar
Manzi, V, Roveri, M, Gennari, R, Bertini, A, Biffi, U, Giunta, S, Iaccarino, SM, Lanci, L, Lugli, S, Negri, A, Riva, A, Rossi, ME and Taviani, M (2007) The deep-water counterpart of the Messinian Lower Evaporites in the Apennine foredeep: the Fanantello section (Northern Apennines, Italy). Palaeogeography, Palaeoclimatology, Palaeoecology 251, 470–99.CrossRefGoogle Scholar
Mosca, P, Polino, R, Rogledi, S and Rossi, M (2010) New data for the kinematic interpretation of the Alps-Apennines junction (Northwestern Italy). International Journal of Earth Sciences 99, 833–49.CrossRefGoogle Scholar
Natalicchio, M, Birgel, D, Peckmann, J, Lozar, F, Carnevale, G, Liu, X, Hinrichs, K-U and Dela Pierre, F (2017) An archaeal biomarker record of paleoenvironmental change across the onset of the Messinian salinity crisis in the absence of evaporites (Piedmont Basin, Italy). Organic Geochemistry 113, 242–53.CrossRefGoogle Scholar
Natalicchio, M, Dela Pierre, F, Birgel, D, Brumsack, H, Carnevale, G, Gennari, R, Gier, S, Lozar, F, Pellegrino, L, Sabino, M, Schnetger, B and Peckmann, J (2019) Paleoenvironmental change in a precession-paced succession across the onset of the Messinian salinity crisis: insight from element geochemistry and molecular fossils. Palaeogeography, Palaeoclimatology, Palaeoecology 518, 4561.CrossRefGoogle Scholar
Natalicchio, M, Dela Pierre, F, Lugli, S, Lowenstein, TK, Feiner, SJ, Ferrando, S, Manzi, V, Roveri, M and Clari, P (2014) Did Late Miocene (Messinian) gypsum precipitate from evaporated marine brines? Insights from the Piedmont Basin (Italy). Geology 42, 179–82.CrossRefGoogle Scholar
Passier, HF, Middelburg, JJ, de Lange, GJ and Böttcher, ME (1997) Pyrite contents, microtextures, and sulfur isotopes in relation to formation of the youngest eastern Mediterranean sapropel. Geology 25, 519–22.2.3.CO;2>CrossRefGoogle Scholar
Pearson, A and Ingalls, AE (2013) Assessing the use of archaeal lipids as marine environmental proxies. Annual Review of Earth and Planetary Sciences 41, 359–84.CrossRefGoogle Scholar
Petrash, DA, Bialik, OM, Bontognali, TRR, Vasconcelos, C, Roberts, JA, McKenzie, JA and Konhauser, KO (2017) Microbially catalyzed dolomite formation: from near-surface to burial. Earth-Science Reviews 171, 558–82.CrossRefGoogle Scholar
Pilskaln, CH and Pike, J (2001) Formation of Holocene sedimentary laminae in the Black Sea and the role of the benthic flocculent layer. Paleoceanography 16, 119.CrossRefGoogle Scholar
Poinsot, J, Schneckenburger, P, Adam, P, Schaeffer, P, Trendel, JM, Riva, A and Albrecht, P (1998) Novel polycyclic sulfides derived from regular polyprenoids in sediments: characterization, distribution, and geochemical significance. Geochimica et Cosmochimica Acta 62, 805–14.CrossRefGoogle Scholar
Rashby, SE, Sessions, AL, Summons, RE and Newman, DK (2007) Biosynthesis of 2-methylbacteriohopanepolyols by an anoxygenic phototroph. Proceedings of the National Academy of Sciences 104, 15099–104.CrossRefGoogle ScholarPubMed
Repeta, DJ and Simpson, DJ (1991) The distribution and recycling of chlorophyll, bacteriochlorophyll and carotenoids in the Black Sea. Deep-Sea Research, Part A. Oceanographic Research Papers 38, S96984.CrossRefGoogle Scholar
Repeta, DJ, Simpson, DJ, Jorgensen, BB and Jannasch, HW (1989) Evidence for anoxygenic photosynthesis from the distribution of bacterio-chlorophylls in the Black Sea. Nature 342, 6972.CrossRefGoogle Scholar
Rosenbaum, J and Sheppard, SMF (1986) An isotopic study of siderites, dolomites and ankerites at high temperatures. Geochimica et Cosmochimica Acta 50, 1147–50.CrossRefGoogle Scholar
Rossi, M, Mosca, P, Polino, R, Rogledi, S and Biffi, U (2009) New outcrop and subsurface data in the Tertiary Piedmont Basin (NW-Italy): unconformity-bounded stratigraphic units and their relationships with basin-modification phases. Rivista Italiana di Paleontologia e Stratigrafia 115, 305–35.Google Scholar
Rouchy, JM and Caruso, A (2006) The Messinian salinity crisis in the Mediterranean basin: a reassessment of the data and an integrated scenario. Sedimentary Geology 188, 3567.CrossRefGoogle Scholar
Roveri, M, Flecker, R, Krijgsman, W, Lofi, J, Lugli, S, Manzi, V, Sierro, FJ, Bertini, A, Camerlenghi, A, De Lange, G, Govers, R, Hilgen, FJ, Hübscher, C, Meijer, PT and Stoica, M (2014) The Messinian Salinity Crisis: past and future of a great challenge for marine sciences. Marine Geology 352, 2558.CrossRefGoogle Scholar
Roveri, M, Lugli, S, Manzi, V and Schreiber, BC (2008) The Messinian salinity crisis: a sequence-stratigraphic approach. GeoActa 1, 169–90.Google Scholar
Ryan, WBF (2009) Decoding the Mediterranean salinity crisis. Sedimentology 56, 95136.CrossRefGoogle Scholar
Sabino, M, Schefuß, E, Natalicchio, M, Dela Pierre, F, Birgel, D, Bortels, D, Schnetger, B and Peckmann, J (2020) Climatic and hydrologic variability in the northern Mediterranean across the onset of the Messinian salinity crisis. Palaeogeography, Palaeoclimatology, Palaeoecology 545, article 109632.CrossRefGoogle Scholar
Santoro, AE, Richter, RA and Dupont, CL (2019) Planktonic marine Archaea. Annual Review of Marine Science 11, 131–58.CrossRefGoogle ScholarPubMed
Sanz-Montero, ME, Rodríguez-Aranda, JP and García Del Cura, MA (2008) Dolomite-silica stromatolites in Miocene lacustrine deposits from the Duero Basin, Spain: the role of organotemplates in the precipitation of dolomite. Sedimentology 55, 729–50.CrossRefGoogle Scholar
Schouten, S, Hopmans, EC, Baas, M, Boumann, H, Standfest, S, Könneke, M, Stahl, DA and Sinninghe Damsté, JS (2008) Intact membrane lipids of ‘Candidatus Nitrosopumilus maritimus,’ a cultivated representative of the cosmopolitan mesophilic group I Crenarchaeota. Applied and Environmental Microbiology 74, 2433–40.Google Scholar
Schouten, S, Hopmans, EC and Sinninghe Damsté, JS (2013) The organic geochemistry of glycerol dialkyl glycerol tetraether lipids: a review. Organic Geochemistry 54, 1961.CrossRefGoogle Scholar
Schouten, S, Pavlović, D, Sinninghe Damsté, JS and de Leeuw, JW (1993) Nickel boride: an improved desulphurizing agent for sulphur-rich geomacromolecules in polar and asphaltene fractions. Organic Geochemistry 20, 901–9.CrossRefGoogle Scholar
Schouten, S, Villanueva, L, Hopmans, EC, van der Meer, MTJ and Sinninghe Damsté, JS (2014) Are Marine Group II Euryarchaeota significant contributors to tetraether lipids in the ocean? Proceedings of the National Academy of Sciences 111, E4285.CrossRefGoogle ScholarPubMed
Sharma, T and Clayton, RN (1965) Measurement of O18/O16 ratios of total oxygen of carbonates. Geochimica et Cosmochimica Acta 29, 1347–53.CrossRefGoogle Scholar
Sierro, FJ, Flores, JA, Francés, G, Vázquez, A, Utrilla, R, Zamarreño, I, Erlenkeuser, H and Barcena, MA (2003) Orbitally-controlled oscillations in planktic communities and cyclic changes in western Mediterranean hydrography during the Messinian. Palaeogeography, Palaeoclimatology, Palaeoecology 190, 289316.CrossRefGoogle Scholar
Sierro, FJ, Flores, JA, Zamarreño, I, Vázquez, A, Utrilla, R, Francés, G, Hilgen, FJ and Krijgsman, W (1999) Messinian pre-evaporite sapropels and procession-induced oscillations in western Mediterranean climate. Marine Geology 153, 137–46.CrossRefGoogle Scholar
Sierro, FJ, Hilgen, FJ, Krijgsman, W and Flores, JA (2001) The Abad composite (SE Spain): a Messinian reference section for the Mediterranean and the APTS. Palaeogeography, Palaeoclimatology, Palaeoecology 168, 141–69.CrossRefGoogle Scholar
Simon, D and Meijer, PT (2017) Salinity stratification of the Mediterranean Sea during the Messinian crisis: a first model analysis. Earth and Planetary Science Letters 479, 366–76.CrossRefGoogle Scholar
Sinninghe Damsté, JS and de Leeuw, JW (1990) Analysis, structure and geochemical significance of organically bound sulfur in the geosphere. State of the art and future research. Organic Geochemistry 16, 1077–101.CrossRefGoogle Scholar
Sinninghe Damsté, JS, Frewin, NL, Kenig, F and de Leeuw, JW (1995a) Molecular indicators for palaeoenvironmental change in a Messinian evaporitic sequence (Vena del Gesso, Italy). I: variations in extractable organic matter of ten cyclically deposited marl beds. Organic Geochemistry 23, 471–83.CrossRefGoogle Scholar
Sinninghe Damsté, JS, Kenig, F, Koopmans, MP, Köster, J, Schouten, S, Hayes, JM and de Leeuw, JW (1995b) Evidence for gammacerane as an indicator of water column stratification. Geochimica et Cosmochimica Acta 59, 1895–900.CrossRefGoogle ScholarPubMed
Sinninghe Damsté, JS, Kuypers, MMM, Schouten, S, Schulte, S and Rullkötter, J (2003) The lycopane/C31 n-alkane ratio as a proxy to assess palaeoxicity during sediment deposition. Earth and Planetary Science Letters 209, 215–26.CrossRefGoogle Scholar
Sinninghe Damsté, JS, Schouten, S, Hopmans, EC, van Duin, ACT and Geenevasen, JAJ (2002) Crenarchaeol: the characteristic core glycerol dibiphytanyl glycerol tetraether membrane lipid of cosmopolitan pelagic crenarchaeota. Journal of Lipid Research 43, 1641–51.CrossRefGoogle Scholar
Sinninghe Damsté, JS, Wakeham, SG, Kohnen, MEL, Hayes, JM and de Leeuw, JW (1993) A 6,000-year sedimentary molecular record of chemocline excursions in the Black Sea. Nature 362, 827–9.CrossRefGoogle ScholarPubMed
Sollai, M, Villanueva, L, Hopmans, EC, Keil, RG and Sinninghe Damsté, JS (2019) Archaeal sources of intact membrane lipid biomarkers in the oxygen deficient zone of the eastern tropical South Pacific. Frontiers in Microbiology 10, article 765.CrossRefGoogle ScholarPubMed
Sturani, C (1973) A fossil eel (Anguilla sp.) from the Messinian of Alba (Tertiary Piedmont Basin). Palaeoenvironmental and palaeogeographic implications. In Messinian Events in the Mediterranean (ed. Drooger, C. W.), pp. 243–55. Amsterdam: Koninklijke Nederlandse Akademie von Wetenschappen.Google Scholar
Sturani, C (1976) Messinian facies in the Piedmont Basin. Memorie Società Geologica Italiana 16, 1125.Google Scholar
Sturani, C and Sampò, M (1973) Il Messiniano inferiore in facies diatomitica nel Bacino Terziario Piemontese. Memorie Società Geologica Italiana 12, 335–38.Google Scholar
Tagliavento, M, Lauridsen, BW and Stemmerik, L (2020) Episodic dysoxia during Late Cretaceous cyclic chalk-marl deposition – Evidence from framboidal pyrite distribution in the upper Maastrichtian Rørdal Mb., Danish Basin. Cretaceous Research 106, article 104223.CrossRefGoogle Scholar
Teixidor, P, Grimait, JO, Pueyo, JJ and Rodriguez-Valera, F (1993) Isopranylglycerol diethers in non-alkaline evaporitic environments. Geochimica et Cosmochimica Acta 57, 4479–89.CrossRefGoogle Scholar
Turich, C and Freeman, KH (2011) Archaeal lipids record paleosalinity in hypersaline systems. Organic Geochemistry 42, 1147–57.Google Scholar
Turich, C, Freeman, KH, Bruns, MA, Conte, M, Jones, AD and Wakeham, SG (2007) Lipids of marine Archaea: patterns and provenance in the water-column and sediments. Geochimica et Cosmochimica Acta 71, 3272–91.CrossRefGoogle Scholar
Van Gemerden, H and Mas, J (1995) Ecology of phototrophic sulfur bacteria. In Anoxygenic Photosynthetic Bacteria (eds Blankenship, RE, Madigan, MT and Bauer, CE), pp. 4985. Dordrecht: Springer.CrossRefGoogle Scholar
van Lith, Y, Warthmann, R, Vasconcelos, C and McKenzie, JA (2003) Sulphate-reducing bacteria induce low-temperature Ca-dolomite and high Mg-calcite formation. Geobiology 1, 71–9.CrossRefGoogle Scholar
Vasiliev, I, Mezger, EM, Lugli, S, Reichart, GJ, Manzi, V and Roveri, M (2017) How dry was the Mediterranean during the Messinian salinity crisis? Palaeogeography, Palaeoclimatology, Palaeoecology 471, 120–33.CrossRefGoogle Scholar
Violanti, D, Lozar, F, Natalicchio, M, Dela Pierre, F, Bernardi, E, Clari, P and Cavagna, S (2013) Stress-tolerant microfossils of a Messinian succession from the Northern Mediterranean basin (Pollenzo section, Piedmont, northwestern Italy). Bollettino della Società Paleontologica Italiana 52, 4554.Google Scholar
Wakeham, SG, Amann, R, Freeman, KH, Hopmans, EC, Jørgensen, BB, Putnam, IF, Schouten, S, Sinninghe Damsté, JS, Talbot, HM and Woebken, D (2007) Microbial ecology of the stratified water column of the Black Sea as revealed by a comprehensive biomarker study. Organic Geochemistry 38, 2070–97.CrossRefGoogle Scholar
Wakeham, SG, Freeman, KH, Pease, TK and Hayes, JM (1993) A photoautotrophic source for lycopane in marine water columns. Geochemica et Cosmochimica Acta 57, 159–65.CrossRefGoogle ScholarPubMed
Wakeham, SG, Sinninghe Damsté, JS, Kohnen, MEL and de Leeuw, JW (1995) Organic sulfur compounds formed during early diagenesis in Black Sea sediments. Geochimica et Cosmochimica Acta 59, 521–33.Google Scholar
Wakeham, SG, Turich, C, Schubotz, F, Podlaska, A, Li, XN, Varela, R, Astor, Y, Sáenz, JP, Rush, D, Sinninghe Damsté, JS, Summons, RE, Scranton, M I, Taylor, GT and Hinrichs, K-U (2012) Biomarkers, chemistry and microbiology show chemoautotrophy in a multilayer chemocline in the Cariaco Basin. Deep-Sea Research Part I: Oceanographic Research Papers 63, 133–56.CrossRefGoogle Scholar
Warren, JK (2010) Evaporites through time: tectonic, climatic and eustatic controls in marine and nonmarine deposits. Earth-Science Reviews 98, 217–68.Google Scholar
Warthmann, R, van Lith, Y, Vasconcelos, C, McKenzie, JA and Karpoff, AM (2000) Bacterially induced dolomite precipitaion in anoxic culture experiments. Geology 28, 1091–4.2.0.CO;2>CrossRefGoogle Scholar
Wilkin, RT, Barnes, HL and Brantley, SL (1996) The size distribution of framboidal pyrite in modern sediments: an indicator of redox conditions. Geochimica et Cosmochimica Acta 60, 3897–912.CrossRefGoogle Scholar
Zeng, Z, Liu, X-L, Farley, KR, Wei, JH, Metcalf, WW, Summons, RE and Welander, PV (2019) GDGT cyclization proteins identify the dominant archaeal sources of tetraether lipids in the ocean. Proceedings of the National Academy of Sciences 116, 22505–11.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1. (a) Distribution of the Messinian evaporites (pink) in the Western Mediterranean Basin and location of the Piedmont Basin (black box; modified from Manzi et al.2013). (b) Structural sketch map of the Piedmont Basin (modified from Bigi et al.1990); the star indicates the location of the Govone section. (c) Schematic profile of the Piedmont Basin, showing the stratigraphic architecture of the Messinian succession; the Govone section (vertical black bar) and the studied interval (red box) are indicated. Note that the gypsum beds are progressively younger towards the depocentre. MSC – Messinian salinity crisis; PB – Piedmont Basin; PLG – Primary Lower Gypsum (modified from Dela Pierre et al.2011).

Figure 1

Fig. 2. Tuning of the Govone section with the astronomical solution (65° N summer insolation; Laskar et al.2004) and correlation with the Perales section (Spain; Sierro et al.2001; Manzi et al.2013). Numbers in circles on the right represent the main bioevents reported in the main text. Bioz. – biozones; FAO – first abundant occurrence (modified from Gennari et al.2020; Sabino et al.2020).

Figure 2

Fig. 3. (a) The Govone section with sample locations, total inorganic carbon contents, carbonate mineralogy and bulk-rock carbon (δ13C) and oxygen (δ18O) stable isotope values. White squares and circles refer to δ13C and δ18O values from Bernardi (2013), whereas black squares and circles refer to values from this study. (b) Outcrop view of the Govone section. The red line corresponds to the onset of the MSC; the white dashed lines indicate the top of marl beds; insets indicate the position of figures (c) and (d). (c) Close-up of pre-MSC laminated organic-rich shales. (d) Close-up of pre-MSC homogenous marls.

Figure 3

Fig. 4. Organic-rich shales. (a) Polished slab cut perpendicular to bedding. Type A and B laminae separated by coarser terrigenous layers (t.l.). The black dashed lines envelope a type A lamina, the white arrows indicate two type B laminae. (b) Fluorescent, lens-shaped faecal pellets (black oval) and flattened, faintly fluorescent aggregate (white arrow); UV-fluorescence image. (c) Irregular aggregates cut parallel to bedding; UV-fluorescence image. (d) Faecal pellet (dashed line) composed of coccoliths; SEM micrograph. (e) Irregular aggregate (dashed line) mostly composed of dolomite microcrystals from the MSC part of the section. The white box indicates the position of (f); SEM micrograph. (f) Globular dolomite microcrystals and scattered pyrite framboids (white arrows); SEM micrograph. (g) Hollow dolomite microcrystals (black arrow and inset) and scattered pyrite framboids (white arrows); SEM micrograph. (h) Planktonic foraminifer test (black arrow) surrounded by coccoliths (white arrows). The inset shows a detail with coccoliths and a small pyrite framboid (arrow); SEM micrograph. (i) Foraminifer test partially filled with pyrite (p); plane-polarized light. The inset is a backscattered electron image of a pyrite infilling, consisting of an aggregate of tiny pyrite framboids. Scale bar in the inset is 20 μm.

Figure 4

Fig. 5. (a) Photomicrographs and (b–f) SEM images of marls. (a) Bioturbation traces in pre-MSC marl (black box); the inset is a close-up image in UV light of the bioturbation traces, highlighted by dashed line. Pyrite: (b) framboid, (c) octahedral crystals. Dolomite microcrystals with (d) cauliflower, (e) globular and (f, arrow) dumbbell-like shapes. The arrows in (e) indicate central hollows in dolomite microcrystals.

Figure 5

Fig. 6. TOC contents, lipid biomarker distributions and caldarchaeol/crenarchaeol ratio (cald/cren ratio) across the onset of the Messinian salinity crisis in the Govone section. Note pentacyclic C30 sulphide (Pentac. sulphide) and isorenieratane (isorenier.) in the tetrahymanol distribution profile and the non-metric scale in the lycopene distribution profile. The red line at the base of marls in cycle Gm30 indicates the onset of the Messinian salinity crisis. GDGT – glycerol dibiphytanyl glycerol tetraethers; Ly – lycopane; THM – tetrahymanol; GDGT-0 – caldarchaeol; GDGT-5 – crenarchaeol.

Figure 6

Table 1. Archaeol, GDGTs (both in % relative to all GDGTs plus archaeol), lycopane and tetrahymanol abundances in pre-MSC and MSC Govone sediments. Max = maximum content, Min = minimum content, GDGT-0 = caldarchaeol, GDGT-5 = crenarchaeol.

Figure 7

Fig. 7. Reconstruction of the water-column structure of the Piedmont Basin at (a) precession minima, insolation maxima and shale deposition and (b) precession maxima, insolation minima and marl deposition during the earliest phase of the Messinian salinity crisis. The black arrows indicate the positions of sections (Ar: Arnulfi; Pz: Pollenzo; Gv: Govone). The thickness of the chemocline is emphasized in (b) to highlight the different conditions described in the text.

Supplementary material: File

Sabino et al. supplementary material

Sabino et al. supplementary material

Download Sabino et al. supplementary material(File)
File 70.7 KB