Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


journal homepage: www.elsevier.com/locate/palaeo

Deep-water paleoenvironmental changes based on early-middle Miocene


benthic foraminifera from Malta Island (Central Mediterranean)
Bianca Russo a, *, Luciana Ferraro b, Cecilia Correggia a, Ines Alberico b, Luca Maria Foresi c,
Mattia Vallefuoco b, Fabrizio Lirer b
a
Dipartimento di Scienze della Terra, dell’Ambiente e delle Risorse - DiSTAR, Università “Federico II”, Complesso Universitario di Monte Sant’Angelo (Edificio 10), Via
Vicinale Cupa Cintia, 21, 80126 Napoli, Italy
b
Istituto di Scienze Marine (ISMAR), CNR Napoli, Calata Porta di Massa, Porto di Napoli, 80133 Napoli, Italy
c
Dipartimento di Scienze Fisiche, della Terra e dell’Ambiente, Università degli Studi di Siena, Via Laterina 8, 53100 Siena, Italy

A R T I C L E I N F O A B S T R A C T

Editor: A Dickson A detailed quantitative, statistical and isotopic study on benthic foraminiferal assemblages from the upper
Burdigalian - lower Langhian of Malta Island (St. Peter’s Pool section) was carried out in order to gain more
Keywords: insights on paleoenvironmental changes that affected the central part of the Mediterranean Basin during the
Early-middle Miocene early-middle Miocene. The most abundant and/or palaecologically significant taxa were grouped into Group A
Benthic foraminifera
(oxic/oligotrophic and oxic/oligo-mesotrophic opportunistic behaving) and Group B (hypoxic/eutrophic) and,
Oxic and hypoxic taxa
on the base of their microhabitat in superficial microhabitat taxa (epifaunal + epifaunal to shallow infaunal
Oxygen and Carbon stable isotopes
Paleoecology taxa), intermediate microhabitat taxa (shallow infaunal + shallow to intermediate infaunal taxa) and deep
Malta microhabitat taxa (deep infaunal species). The changes in the structure of the benthic foraminiferal assemblages,
Central Mediterranean Basin together with isotopic data and statistical analysis results, identify five main intervals SPP1-SPP5, testifying the
evolution of bottom water conditions between 16.12 and 15.36 Ma, corresponding to warm climate conditions
due to the onset of the Miocene Climatic Optimum (MCO). The main benthic foraminiferal turnover occurred
between 16.01 and 15.91 Ma. It has been recorded in both the palaecological and geochemical data and can be
related to the Miller’s Mi2 cooling event. This turnover correlates to a sea level falling that we believe caused an
increase of the primary productivity due to runoff intensification. We suggest that the fluctuations of superficial
microhabitat oxic/oligo/oligo-mesotrophic Group A, opposite to those of intermediate and deep microhabitat
hypoxic/eutrophic Group B, and the occurrence of the superficial microhabitat, oxic/oligo-mesothrophic
opportunistic behaving species are indicative of increased seasonality, which in turn affect food flux and
oxygenation at the sea-floor. Moreover, the calculation of %P and the well-defined depth distribution range of
some significant species allowed us to estimate a middle bathyal paleodepth, ranging from 600 m to 1000 m for
the entire composite section.

1. Introduction and fauna evolved into the same taxa that exist today. Miocene climate
was dynamic: long periods of early and late glaciation bracketed a ~ 2
The global oceanographic system, from the Indo-Pacific to Atlantic Myr greenhouse interval, the Miocene Climatic Optimum (MCO). Floras,
through the Tethys Ocean, was affected by a complex and dynamic faunas, ice sheets, precipitation, pCO2, and ocean and atmospheric cir­
climatic history mostly due to ocean reorganization during the Late culation mostly (but not ubiquitously) covaried with these large changes
Oligocene-Miocene time interval (Miller et al., 1991; Zachos et al., 2001, in climate. With higher temperatures and moderately higher pCO2
2008). (~400–600 ppm), the MCO has been suggested as a particularly
The Miocene epoch (23.03–5.33 Ma) was a time interval of global appropriate analog for future climate scenarios, and for assessing the
warmth, relative to today. Continental configurations and mountain predictive accuracy of numerical climate models, the same models that
topography transitioned towards modern conditions, and many flora are used to simulate future climate (Steinthorsdottir et al., 2021). The

* Corresponding author.
E-mail address: bianca.russo@unina.it (B. Russo).

https://doi.org/10.1016/j.palaeo.2021.110722
Received 23 January 2021; Received in revised form 7 October 2021; Accepted 15 October 2021
Available online 30 October 2021
0031-0182/© 2021 Elsevier B.V. All rights reserved.
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

middle Miocene represents an important time interval for the climate benthic foraminifera and of the benthic δ18OH. dutemplei and δ13CH. dutemplei
evolution (i.e., Shackleton and Kennett, 1975; Woodruff and Savin, isotope records of the St. Peter’s Pool section (Malta Island, Central
1989, 1991; Holbourn et al., 2005; Verducci et al., 2009), which grad­ Mediterranean) are presented.
ually moved from warm MCO, termed by the Middle Miocene Climate
Transition (MMCT), to cooler conditions due to major environmental 2. Study area: The Maltese Archipelago
changes (Miller and Kent, 1987; Miller et al., 1991; Zachos et al., 2001),
mostly driven by orbital forcing (Holbourn et al., 2007, 2013b, 2015). 2.1. Geological setting
During this period, δ18O data on open ocean benthic foraminifera reveal
three prominent enrichments that reflect increased glaciation, known as The Maltese Archipelago is located in the Central Mediterranean,
Mi2a, Mi3a and Mi3b Miller’s events (Miller et al., 1991). In particular, approximately 80 km South of Sicily and 320 km North of North Africa
the Mi3b event represents the most prominent δ18O excursion associated (Fig. 1), close to the actual boundary between the Malta Platform and
with the major expansion of the East Antarctic Ice Sheet at ~13.8 Ma, in the Pantelleria Rift (Finetti, 1982).
the uppermost part of MMCT interval (Miller et al., 1991). In addition, a The archipelago extends for 45 km in a NW-SE direction and consists
long-lasting positive excursion in δ13C benthic signal (“Monterey of 3 main islands: Malta, Gozo and Comino (Fig. 1), and several small
Excursion” between 17 and 13.5 Ma; Zachos et al., 2001) is documented uninhabited islets which include Cominotto, Filfola, St. Paul’s Islands,
from Langhian deep ocean (i.e., Holbourn et al., 2007, 2013a,b, 2015 Fungus Rock and few other minor rocks.
and reference therein) to shallow water Mediterranean records (Bran­ According to the geological literature (Hyde, 1955; Giannelli and
dano et al., 2021 and reference therein). Salvatorini, 1972, 1975; Felix, 1973; Pedley, 1976, 1978; Pedley et al.,
In terms of Miocene proto-Mediterranean, significant paleogeo­ 1976, 1978; Mazzei, 1985; Föllmi et al., 2008; Foresi et al., 2008;
graphic and paleoenvironmental changes interested this area. In Gruszczynski et al., 2008; Baldassini et al., 2013; Foresi et al., 2014;
particular, the counter-clockwise rotation of Africa and Arabia caused a Baldassini and Di Stefano, 2017), the stratigraphy of the Maltese Ar­
collision with the Anatolian plate during the late Burdigalian. Then, in chipelago has been exhaustively investigated thanks to the extensive
the early-middle Miocene, the Indo-Pacific connection closed and the and well-exposed sedimentary successions. Sedimentary sequences
Tethyan Seaway transformed in a series of semi-enclosed basins similar include five formations characterized by different lithological, sedi­
to the present-day situation (Rögl and Steininger, 1983; Rehault et al., mentological and paleontological features: 1) the Lower Coralline
1985; Steininger et al., 1985; Woodruff and Savin, 1989; Ramsay et al., Limestone Formation (LCLf) - late Oligocene; 2) the Globigerina Lime­
1998; Harzhauser et al., 2007; Karami et al., 2011; De La Vara and stone Formation (GLf) - Chattian-Langhian, subdivided into three
Meijer, 2016). members, the Lower (LGLm), Middle (MGLm) and Upper Globigerina
According to this paleoclimatic and paleoceanographic framework, Limestone (UGLm) (Rizzo, 1932); 3) the Blue Clay Formation (BCf) -
the upper Burdigalian-Langhian St. Peter’s Pool section in Malta Island Serravallian-Tortonian; 4) the Greensand Formation (GSf), early Mes­
(Foresi et al., 2011) is a key area to analyse the deep-water paleo­ sinian, and 5) the Upper Coralline Limestone Formation (UCLf) - pre-
environmental changes over the MCO interval. Moreover, using benthic evaporitic, early Messinian.
foraminifera and stable isotope signatures, δ18O excursion related to
Mi2a event will be identified for the first time.
Benthic foraminifera are useful tool to detect and monitor the 2.2. St. Peter’s pool section
modern and past deep water-mass ocean circulation (e.g. Bernhard and
Sen Gupta, 1999; Jorissen, 1999; Loubere and Fariduddin, 1999; Kou­ The St. Peter’s Pool section outcrops on the eastern coast of the
wenhoven and van der Zwaan, 2006; Sarkar and Gupta, 2009; Mourik Delimara Peninsula (Fig. 1) and is representative of the Upper Globi­
et al., 2011; Eichler et al., 2016; Rathburn et al., 2018; Xue et al., 2019) gerina Limestone member (Foresi et al., 2011).
and relationship with chemical and physical features of deep water The section is 31 m thick and lies on the C2 phosphate-rich bed
masses was highlighted by several authors (e.g. Smart and Ramsay, (Fig. 2).
1995; Schmiedl et al., 1997; Yasuda, 1997; Bellanca et al., 2002; Spro­ The sedimentary record is characterized by a cyclic alternation of
vieri et al., 2004; Jorissen et al., 2007; Russo et al., 2007; Pérez-Asensio calcareous marl, marly limestone and jutting bioturbated hardened
et al., 2012). Benthic foraminifera are closely related to oxygen con­ limestone. In the middle part of the section a thick dark-grey interval,
centration in the bottom waters and in the sediment, as well as to the recognizable in the whole Delimara Peninsula, occurs between 10.2 and
organic matter input from the surface waters (e.g. Moodley and Hess, 11.9 m (Fig. 2). The lower part of the section, from the C2 phosphate-
1992; Sen Gupta and Machain-Castillo, 1993; Gooday, 1994; Kaiho, rich bed to the dark-grey interval, consists of light beige coloured
1994; van der Zwaan et al., 1999; Sander Ernst and van der Zwaan, rhythmic alternations of calcareous marl, marly limestone and dark
2004; Murray, 2006; Pucci et al., 2009; Caulle et al., 2013, 2015; Drinia beige hardened limestone that are rich in echinoid rests, molluscs and
et al., 2018). fossil tracks. Conversely, the upper part of the section, from the dark-
Previous studies on Miocene benthic foraminiferal assemblages from grey interval to the top, is characterized by a marked change in colour
Maltese Archipelago were carried out by Bellanca et al. (2002) on Ras il- from light beige to dark beige, which is associated to a progressive
Pellegrin section. The authors provide significant results obtained from decrease of the calcareous marl thickness and to an equivalent increase
isotope and benthic foraminiferal data, suggesting a direct relationship in prominent hardened limestone. At 28.4 m, calcareous marl and marly
between surface and intermediate Mediterranean water masses during limestone change colour again and become whitish (Fig. 2). For a more
the Langhian-Serravallian. Ostracods from the Langhian-Serravallian of detailed description of the section, see Foresi et al. (2011).
Ras il-Pellegrin composed section were studied by Bonaduce and Barra The integrated bio-magnetostratigraphic study proposed by Foresi
(2002) to define changes in palaeoenvironmental conditions at the sea- et al. (2011) allowed the authors to attribute the section to the upper­
floor. Benthic foraminifera from St. Peter’s Pool section were roughly most Burdigalian to lower Langhian, spanning from planktonic forami­
analyzed by Foresi et al. (2011), in order to estimate the paleodepth of nifera biozones MMi3 to MMi4c pp. (Foresi et al., 2011; Lirer et al.,
the marine sedimentary record. 2019) and the calcareous nannofossil zones MNN4a to MNN5a (Foresi
In this paper the benthic foraminifera together with Oxygen and et al., 2011) (Fig. 2). The age model for S. Peter’s Pool section bases on
Carbon stable isotopes have been used as proxy to evaluate the main calcareous plankton biovents by Foresi et al. (2011) and astronomical
palaeoecological changes of the bottom water masses occurred in the ages proposed by Turco et al. (2017) in the La Vedova section (Conero
proto-Mediterranean during the upper Burdigalian - lower Langhian. Area, Central Italy). The age-depth profile has been produced by linear
The results obtained from the quantitative and statistical study of interpolation between the identified biostratigraphic tie-points (Fig. 2).

2
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

Fig. 1. Location and geological map of the study area (modified from Foresi et al., 2011); geology of Delimara Peninsula from Oil Exploration Directorate (1993).

3. Material and methods 2003, 2006). Moreover there is a general consensus that oligotrophic
and highly oxygenated environments are relatively enriched in super­
3.1. Sample preparation and classification of benthic foraminifera ficially living taxa, whereas eutrophic and oxygen limited environments
are relatively dominated by deep infaunal species (Sgarrella et al.,
A total of 135 samples for benthic foraminiferal analyses were 2012). The motile benthic deep-water foraminifera have the capability
collected from the investigated section (Appendix A) and processed to rapidly adapt their microhabitat to changes in food availability and
using standard laboratory procedures. Specifically, they were dis­ oxygenation. This dynamic adaptation is also reflected in the so-called
aggregated in normal water, washed and sieved into two size fractions TROX model (TROX = TRophic OXygen model, Jorissen et al., 1995,
(mesh sizes 63 μm and 125 μm). The fraction >125 μm was splitted by an 2007), which explains that the depth of the foraminiferal microhabitat is
Otto microsplitter to obtain a representative fraction, containing controlled by food availability in oligotrophic ecosystems and oxygen
approximately 300 benthic foraminiferal specimens. All specimens were concentration in eutrophic ecosystems. In oligotrophic environments,
picked out, identified at specific or supraspecific level, counted and the microhabitat depth is limited by the low amount of food available
stored in Chapman slides. The taxonomic identification was based for within the sediment, whereas in eutrophic systems, the penetration
genera on Loeblich and Tappan (1964, 1988), and for species on Van depth of most taxa depends on the level of oxygen present in the sedi­
Morkhoven et al. (1986), Holbourn et al. (2013a). Some online data­ ment (Jorissen et al., 1995).
bases, such as Foraminifera.eu (https://foraminifera.eu/) and Worms In Table 1 the ecological and microhabitat preferences of the
(https://www.marinespecies.org/; Hayward et al., 2018) have also been recognized benthic taxa are listed and the relative references are
used. reported.

3.2. Ecological preferences of benthic foraminiferal taxa 3.3. Proxies based on benthic foraminifera

According to Sgarrella et al. (2012) Recent deep-sea benthic fora­ Generally, foraminifera are abundant and their preservation is quite
minifera are dependent on food availability, in terms of productivity good. Then, it was possible to obtain:
exported to the sea floor, essential to meet their energy requirements
(Loubere and Fariduddin, 1999), and on oxygen bottom water content - relative frequencies (%) of the most abundant and/or palae­
(e.g. Corliss, 1985; Gooday, 1988, 1993; Herguera and Berger, 1991; cologically significant taxa to describe the main paleoenvironmental
Barmawidjaja et al., 1992; Jorissen et al., 1995; Jorissen and Wittling, changes at the sea-floor;
1999; De Rijk et al., 2000; Schmiedl et al., 2000; Fontanier et al., 2002,

3
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

2017) were subdivided into three groups basing on their different


sensibility to oxygen content at the sea floor. The three groups
correspond to three increasing oxygen levels: low (Group 1), inter­
mediate (Group 2), and high (Group 3) level of oxygen bottom
waters.
- plankton/benthos ratio, expressed as the percentage of planktonic
specimens, %P = 100*[P/(P + B-S)], where P = number of plank­
tonic specimens, B = number of benthic specimens and S = number
of stress markers (deep infauna) (van Hinsbergen et al., 2005). This
proxy was used for paleodepth estimation. Bolivina spp., Bulimina
elongata, Cancris spp., Globobulimina affinis, Uvigerina spp. (except for
U. semiornata) have been considered as stress markers following van
Hinsbergen et al., 2005. The paleodepth estimation obtained by this
proxy was confirmed and refined by means of the co-occurrence of
depth benthic foraminiferal markers with well-defined depth distri­
bution range.

3.4. Stable isotopes

Oxygen and Carbon stable isotopes analyses were carried out to


describe the main changes of temperature and productivity of bottom
water masses. In particular, 283 samples disaggregated in normal water,
no oven dried, were analyzed to pick about 10 specimens of the benthic
foraminifera species Heterolepa dutemplei, considered indicative of oxic
bottom conditions (e.g. Russo et al., 2007). For our purpose clean in­
dividuals without concretions, overgrowth or visible diagenetic pro­
cesses on going were selected from grain size fraction ≥150 μm.
Analyses were performed at the geochemistry laboratory of the
ISMAR-CNR (Naples, Italy) with an automated continuous flow car­
Fig. 2. The St. Peter’s Pool section: Polarity/Chron, lithology, planktonic bonate preparation Gas BenchII device (Spötl and Vennemann, 2003)
foraminifera and calcareous nannoplancton bioevents, and biostratigraphy and a ThermoElectron Delta Plus XP mass spectrometer. Acidification of
(from Foresi et al., 2011). Age-depth profile is based on calcareous plankton samples was performed at 50 ◦ C. Every six samples, an internal standard
ages of Turco et al. (2017). For samples references see Appendix A. (Carrara Marble with δ18O = − 2.43‰ versus VPDB and δ13C = 2.43‰ vs.
VPDB) was used. Moreover, every 30 samples the NBS19 international
standard was measured. Standard deviations of carbon and oxygen
- relationship between taxa of Group A (oxic/oligotrophic taxa and isotope measures were estimated at 0.1 and 0.08‰, respectively, basing
oxic/oligo-mesotrophic opportunistic behaving species) and Group B on three repetitions of ~200 samples measures. All the isotope data are
(hypoxic/eutrophic species) (Table 2) to reconstruct the main reported in δ‰ vs. VPDB.
palaecological changes. We use the term hypoxic in analogy to sensu
Jorissen et al. (2007) for the environments, referring to the species 3.5. Statistical analysis
that prefer environments where they may potentially be influenced
by low oxygen concentrations (without referring to a precise range of A multivariate statistical analysis method was applied (R-mode
oxygen concentration). For literature references related to the PCA), using SPSS statistical software (version 19) to quantify the rela­
ecological characteristics of each taxon see section 3.2, Table 1; tionship between benthic foraminiferal species recognized in 135 sam­
- relative proportions of superficial microhabitat taxa (epifaunal to ples and evaluate the major changes in the benthic foraminiferal
shallow infaunal + epifaunal taxa), intermediate microhabitat taxa assemblages. Starting from the original variables set, the PCA generates
(shallow to intermediate infaunal + shallow infaunal taxa), and deep a new set of uncorrelated variables named Principal Components (PCs),
infaunal microhabitat taxa. Hanzawaia boueana was excluded from resulting from a linear combination of the original ones. The PCs explain
the group of the superficial microhabitat taxa, because even if the percentage of the system’s total variability concentrated on the first
epifaunal species (Spezzaferri et al., 2014), it is reported as tolerant two or three main components (Davis, 1973). Statistical analysis was
to high food supply and/or great preservation of organic matter at applied to the 20 taxa with an abundance >4%, out of 42 forming the
the sea floor (Russo et al., 2002) and indicative of high productivity original dataset, moreover other 3 species (Anomalinoides helicinus,
(Russo et al., 2007); Cassidulina laevigata, Globocassidulina subglobosa) with Pearson correla­
∑ ∑ tion lower than 0.3 were not considered.
- paleoxygenation index - O2 index [O2 index = (1 (1) + 2 (2) +
∑ ∑
3 (3) + …. N (n))/100] where Σ(n) is the sum of the relative 4. Results
abundances of taxa in group n, following Kouwenhoven and van der
Zwaan (2006). This proxy is useful to detect the trend of the oxygen 4.1. Benthic taxa distribution patterns
level at the sea floor. For this purpose, the benthic taxa with per­
centages >4% (with the exception of Melonis affinis because we Benthic foraminiferal assemblages are composed of 42 taxa, referred
considered it poorly affected by the oxygen content and controlled to the genera Anomalinoides, Bolivina, Bulimina, Cancris, Cassidulina,
more by temperature than food supply - Rasmussen and Thomsen, Cibicides, Cibicidoides, Cribroelphidium, Epistominella, Globobulimina,

4
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

Table 1
Ecological and microhabitat preferences of the studied benthic taxa.
Species Microhabitat Ecological preference

Anomalinoides helicinus Epifaunal/shallow infaunal (Pérez-Asensio et al., 2012). Species generally preferring oxic and/or oligotrophic environments (
Epifaunal (Sgarrella et al., 2012). Sgarrella et al., 2012).
Bolivina reticulata Shallow infaunal (Pérez-Asensio et al., 2012). Species belonging to Bolivina spp., reported in Székely et al. (2017) as low-
oxygen group (Spezzaferri et al., 2002), which includes some species
tolerating dysoxia (Murray, 2006); dysoxic species (Kaiho, 1994).
Bolivina spathulata Deep infaunal (Violanti et al., 2011). Shallow infaunal (Singh et al., Species abundant in poor oxygen, rich in organic matter with elevated
2015, Pascual et al., 2020). productivity environments (Alavi, 1988; Martin et al., 2007; Singh et al.,
2015; Pascual et al., 2020). Species behaving as opportunistic when a great
amount of organic matter is available, and usually associated to upwelling
phenomena (Okada, 1983; Mathieu, 1986; Guerreiro et al., 2009).
Species indicating dysoxia (Kaiho, 1994; Pascual et al., 2020).
Bulimina striata Shallow infaunal (Grunert et al., 2018). Species indicative of high productivity (Bellanca et al., 2002; Russo et al.,
2007).
Species tolerating dysoxia (Murray, 2006; Székely et al., 2017), and low
oxygen conditions (Spezzaferri et al., 2002; Székely et al., 2017) since
belonging to Bulimina spp.
Species indicative of OMZ of eastern Arabian Sea, can be used as an index
for delineating oxygen-depleted environments in Arabian Sea (Nigam et al.,
2007).
Cassidulina laevigata Shallow infaunal (Fontanier et al., 2002, Pascual et al., 2020). Opportunistic species adapted to high food concentration (Baas et al.,
1998; Pascual et al., 2020), related to high flux of organic carbon (Jorissen
et al., 2007; Singh et al., 2015; Pascual et al., 2020). Suboxic species (
Kaiho, 1994), related with upwelling currents (Levy et al., 1995; Pascual
et al., 2020). C. laevigata dominates the Younger Dryas (YD) and the
beginning of Lower Holocene, evidencing the occurrence of cool waters
with low-oxygen and high organic matter content (Pascual et al., 2020).
High stress, dysoxia and high organic matter tolerant species (Rögl and
Spezzaferri, 2003; Murray, 2003, 2006; Alve, 2010; Spezzaferri et al.,
2013).
C. laevigata group (including C. laevigata and Trifarina angulosa): prefers
warm and food rich conditions (Rasmussen and Thomsen, 2017). Species
belonging to Group 1, comprising Assemblages 1 and 2, which include
species that dominated during warm and mostly food-rich interstadial
warm phases and interstadial cooling phases (Rasmussen and Thomsen,
2017).
Cibicides pachyderma Epifaunal to shallow infaunal (Pérez-Asensio et al., 2012; Spezzaferri Oxic/oligo-mesotrophic species (Fontanier et al., 2002; Sgarrella et al.,
et al., 2014). 2012; Spezzaferri et al., 2013; Pascual et al., 2020); opportunistic species
characteristic of a varying organic supply in relatively oligotrophic or
oligo-mesotrophic settings (Schmiedl et al., 2003; Abu-Zied et al., 2008;
Melki et al., 2009; Sgarrella et al., 2012); species able to compete in an
unstable environment (Sgarrella et al., 2012) and occurring in upwelling
areas (Licari and Mackensen, 2005).
Cibicidoides cicatricosus Epifaunal (Murray, 1991). Oxic species (Kaiho, 1994; Russo et al., 2007).
Cibicidoides subhaidingerii/ Epifaunal (Murray, 2006; Székely et al., 2017). C. subhaidingerii: oxic (Corliss, 1991; Kaiho, 1994; Oblak, 2007; Russo
Heterolepa dutemplei et al., 2007; Sgarrella et al., 2012; Székely et al., 2017) and oligotrophic
species, since belongs to Cibicidoides spp. generally preferring oxic and/or
oligotrophic environments (Sgarrella et al., 2012).
Cibicidoides ungerianus Epifaunal (Oblak, 2007). Epifaunal to shallow infaunal (Pérez-Asensio Species indicating a rather oligotrophic environment and oxic bottom
et al., 2012). conditions (Sgarrella et al., 2012).
Globocassidulina subglobosa Epifaunal to shallow infaunal (Pérez-Asensio et al., 2012). G. subglobosa assemblage seems to be adapted to an environment that is
characterized by low organic matter fluxes and enhanced bottom current
velocities (Schmiedl et al., 1997). Oxic indicator (Kaiho, 1999). Species
indicating relatively nutrient-poor conditions and recorded in areas of low
seasonal food supply in the North Atlantic and reported as opportunistic
species (Rasmussen et al., 2002). Species generally preferring oxic and/or
oligotrophic environments (Sgarrella et al., 2012). Oligotrophic behaving (
Spezzaferri et al., 2013).
Species with an intermediate sensibility to the presence of oxygen at the sea
floor (group 3 of a scale going from 1 -low-oxygen conditions- to 5 -normal
to well oxygenated conditions-) (Kouwenhoven and Van der Zwaan, 2006).
Gyroidina spp. Shallow to intermediate infaunal (Fentimen et al., 2018). Suboxic indicator (Kaiho, 1999).
Hanzawaia boueana Epifaunal (Spezzaferri et al., 2014). Species tolerant to high food supply and/or great preservation of organic
matter at the sea floor (Russo et al., 2002), and indicative of high
productivity (Russo et al., 2007).
Species with a relative low sensibility to the presence of oxygen at the sea
floor (group 2 of a scale going from 1 -low-oxygen conditions- to 5 -normal
to well oxygenated conditions-) (Kouwenhoven and Van der Zwaan, 2006).
Lenticulina spp. Epifaunal (Di Bella et al., 2015; Székely et al., 2017). Group reported in Székely et al. (2017) as suboxic (Kaiho, 1994; Rögl and
Spezzaferri, 2003; De Man, 2006; Grunert et al., 2012), and cold (Murray,
2006).
Melonis affinis Shallow to intermediate infaunal (Fontanier et al., 2002; Grunert et al., Cold water species (− 0.4◦ to 9 ◦ C) (Murray, 1991; Pascual et al., 2020),
2015; Singh et al., 2015; Pascual et al., 2020). tolerating high degraded organic matter content (Fontanier et al., 2002;
Pascual et al., 2020) and indicating mesotrophic conditions (Jorissen,
2003; Grunert et al., 2015; Pascual et al., 2020). The species is attracted to
(continued on next page)

5
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

Table 1 (continued )
Species Microhabitat Ecological preference

colder water and low food supply as included in Group 2 species that
dominated during cold and generally food poor interstadial cold phases and
stadials (Rasmussen and Thomsen, 2017). Species controlled more by
temperature than food supply (Rasmussen and Thomsen, 2017). Suboxic
indicator (Kaiho, 1994, 1999), it can occur in oxygen depleted sediments (
Jorissen et al., 1998; Pascual et al., 2020).
Oridorsalis umbonatus Epifaunal to shallow infaunal (Pérez-Asensio et al., 2012). Suboxic indicator since belongs to Oridorsalis spp. (Kaiho, 1999).
Pullenia bulloides Shallow infaunal (Pérez-Asensio et al., 2012). Suboxic indicator since belongs to Pullenia spp. (Kaiho, 1999). The species
prefers colder water and low food supply as included in Group 2 species
that dominated during cold and generally food poor interstadial cold
phases and stadials (Rasmussen and Thomsen, 2017). Species controlled
more by temperature than food supply (Rasmussen and Thomsen, 2017).
Siphonina reticulata Epifaunal to shallow infaunal (Pérez-Asensio et al., 2012). Oxic species (Kouwenhoven and van der Zwaan, 2006; Russo et al., 2007;
Sgarrella et al., 2012; Pérez-Asensio et al., 2020), generally preferring
oligotrophic environments (Sgarrella et al., 2012).
Spirorutilus carinatus Epifaunal to shallow infaunal (Pérez-Asensio et al., 2012). High oxic conditions species (Ied et al., 2011).
Uvigerina cylindrica Shallow infaunal (Stefanelli, 2004). Species proliferating during periods of nutrient abundance (stressed and
gaudrynoides nutrient-rich environments) (Drinia, 2018). Suboxic species (Kaiho, 1994;
Rögl and Spezzaferri, 2003; Székely et al., 2017), indicative of high organic
matter (Rögl and Spezzaferri, 2003; Székely et al., 2017), and high primary
productivity (Spezzaferri et al., 2002; Székely et al., 2017), as belonging to
Uvigerina spp.
Uvigerina peregrina Shallow infaunal (Pérez-Asensio et al., 2012; Spezzaferri et al., 2014; Species occuring in sediments rich in organic matter (Seiglie, 1968;
Milker et al., 2019). Shallow to intermediate infaunal (Fontanier et al., Schmiedl et al., 1997; Fontanier et al., 2002; Singh et al., 2015; Milker
2002; Singh et al., 2015; Pascual et al., 2020). et al., 2019; Pascual et al., 2020), indicating high productivity (Lutze,
1986; Rai and Singh, 2012; Patarroyo-Camargo and Martinez-Rodriguez,
2013; Pascual et al., 2020), meso-to eutrophic conditions (Milker et al.,
2019).
Suboxic species (Kaiho, 1994; Lutze, 1986; Singh et al., 2015; Pascual
et al., 2020).

Globocassidulina, Hanzawaia, Karreriella, Melonis, Neoeponides, Ori­ Ma) up to 15.91 Ma, when most taxa of Group A abruptly decrease
dorsalis, Pullenia, Saracenaria, Sigmoilina, Sigmoilopsis, Siphonina, starting to increase again at 15.92 Ma (Fig. 4c). By contrast, in the same
Sphaeroidina, Spiroplectinella, Trifarina, Uvigerina, and by other three interval Group B abruptly increases starting to decrease at 15.92 Ma
taxonomic units, such as Cibicidoides subhaidingerii/Heterolepa dutemplei, (Fig. 4d). U. cylindrica gaudrynoides occurs almost only in this interval
Gyroidina spp. and Lenticulina spp. The list of the identified taxa is re­ (maximum value 11%), C. laevigata is present with some of its highest
ported in Appendix B. percentage values (maximum value 28%) (Fig. 3).
The abundance fluctuations of the 17 species and of the taxon Cibi­ From 15.91 to 15.86 Ma, the Group A increases (Fig. 4a), while the
cidoides subhaidingerii/Heterolepa dutemplei with percentage > 4% are Group B decreases (Fig. 4b). Conversely from 15.86 to 15.83 Ma, the
reported in Fig. 3. trends of these two groups are mainly reversed (Fig. 4a, b). From 15.83
Throughout the studied section Bolivina spathulata (0.4–50%), Cas­ to 15.71 Ma, the Group A increases or shows positive peaks (Fig. 4a)
sidulina laevigata (0.3–27.5%), C. subhaidingerii/H. dutemplei while the Group B decreases (Fig. 4b). U. cylindrica gaudrynoides disap­
(0.7–39.8%), Siphonina reticulata (0.3–26.5%) and Uvigerina peregrina pears at 15.78 Ma (Fig. 3). From 15.71 to 15.62 Ma the Group A de­
(1.3–30.8%) are dominant species, together with the subordinate Buli­ creases (Fig. 4a) and the Group B exhibits either an increasing trend or
mina striata, Cibicides pachyderma, Cibicidoides cicatricosus, Cibicidoides positive peaks (Fig. 4b). Oxic/oligo-mesotrophic opportunistic behaving
ungerianus, Hanzawaia boueana and Melonis affinis. Further important species (C. pachyderma, C. cicatricosus and C. ungerianus), which
species include Anomalinoides helicinus, Spirorutilus carinatus and Uvi­ contribute to the fluctuating trend of Group A from 15.91 to 15.62 Ma,
gerina cylindrica gaudrynoides. occur with the highest percentage values up to 15.71 Ma (Fig. 4d). From
In the distribution patterns described below we grouped the most 15.62 to 15.37 Ma the Group A increases (Fig. 4a), while the Group B
abundant and subordinate taxa, and the other important species such as shows an overall slightly decreasing trend (Fig. 4b), and the oxic/oligo-
Anomalinoides helicinus, Spirorutilus carinatus and Uvigerina cylindrica mesotrophic/opportunistic behaving species occur with a discontinuous
gaudrynoides into Group A (A. helicinus, C. subhaidingerii/H. dutemplei, trend (Fig. 4d). B. spathulata shows a decreasing trend, even if it occurs
C. cicatricosus, C. pachyderma, C. ungerianus, S. carinatus, S. reticulata) with percentage values up to 33% from 15.55 to 15.50 Ma (Fig. 3). From
and Group B (B. striata, B. spathulata, C. laevigata, U. cylindrica gau­ 15.37 to 15.36 Ma, the Group A moderately decreases (Fig. 4a), while
drynoides, U. peregrina) on the base of their ecological preferences the Group B slightly increases (Fig. 4b) and the oxic/oligo-mesotrophic/
(Table 2). The description of the patterns is mainly referred to the two opportunistic behaving species drastically reduce (Fig. 4d), and
Groups together with the trend of some taxa which we considered B. spathulata increases up to 50% at 15.37 Ma (Fig. 3). C. laevigata and
particularly significant for their ecological meaning. M. affinis are quite abundant throughout the section and both show a
From 16.12 to 16.01 Ma, Group A is dominant with respect to Group significant increasing in abundance from 16.08 to 16.06 Ma and from
B (Fig. 4c, d). C. subhaidingerii/H. dutemplei displays abundance peaks 16.01 to 15.91 Ma. Upwards C. laevigata has a decreasing trend up to
(40%, 29%) at 16.08 and 16.03 Ma; B. spathulata shows its lowest per­ 15.64 Ma showing high abundances percentage values up to 16.62 Ma,
centage values starting from 16.08 Ma, B. striata displays an abundance and then decrease up to 15.54 Ma to increase again up to 15.37 Ma,
peak (34%) at 16.11 Ma, C. laevigata has three peaks (26%, 20%, 22%) at where decreases up to the top of the section; M. affinis shows an
16.12, 16.07 and 16.01 Ma, U. cylindrica gaudrynoides displays a peak increasing trend from 15.78 to 15.60 Ma, where its highest percentage
(16%) at 16.09 Ma, and U. peregrina shows high percentage values values (about 10%) are recorded (Fig. 3), from 15.60 to 15.50 Ma its
(maximum value 20%) from 16.05 to 16.03 Ma (Fig. 3). Substantial percentage abundance values decrease to slightly increase again up to
changes in the assemblage abundance and composition are observed the top of the succession. H. boueana is present along the whole section
starting from the transitional thick dark-grey interval (16.01 to 15.96 with low values (Fig. 3) and it shows an opposite trend to oxic/

6
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

Table 2
List of oxic/oligotrophic and oxic/oligo-mesotrophic opportunistic behaving taxa (Group A), and hypoxic/eutrophic taxa (Group B). The species C. cicatricosus
and C. ungerianus have been considered oxic/oligo-mesotrophic opportunistic behaving species as C. pachyderma, since they mirror the distribution pattern of
the last species. The taxa are also grouped on the base of their preferred microhabitat: superficial microhabitat taxa (epifaunal + epifaunal to shallow
infaunal), intermediate microhabitat taxa (shallow infaunal + shallow to intermediate infaunal) and deep microhabitat taxa (deep infaunal). The taxa below
the thick black line are not included in Group A and Group B because: Lenticulina spp. is an epifaunal taxa but reported as suboxic group; M. affinis is poorly
affected by the oxygen content and controlled more by temperature than food supply; the other taxa do not show any significant change along the section. For
references see section 3.2 Table 1.
Group A Group B

Oxic/oligo-mesotrophic
Oxic/oligotrophic taxa Hypoxic/eutrophic species
opportunistic behaving species

Deep microhabitat
Superficial microhabitat taxa Intermediate microhabitat taxa
taxa
Epifaunal Epifaunal to Epifaunal to shallow Shallow to
shallow Epifaunal infaunal Shallow infaunal intermediate Deep infaunal
infaunal infaunal
Anomalinoides
Bulimina striata
helicinus
Cibicidoides Cibicides pachyderma
Siphonina Cassidulina
subhaidingerii/ Cibicidoides
reticulata laevigata Bolivina spathulata
Heterolepa cicatricosus
Spirorutilus Uvigerina cylindrica
dutemplei Cibicidoides
carinatus gaudrynoides
ungerianus
Uvigerina peregrina
Globocassidulina
Bolivina reticulata Gyroidina spp.
Lenticulina subglobosa
spp. Oridorsalis
Pullenia bulloides Melonis affinis
umbonatus

oligotrophic taxa (Fig. 4c). taxa slightly decrease while intermediate and deep microhabitats taxa
increase.
4.2. Benthic foraminiferal microhabitats
4.3. Statistical analyses
In Table 2 the benthic taxa with percentages >4% are also grouped
on the base of their microhabitat (for references see section 3.2, Table 1). The selected benthic taxa show a good level of correlation confirmed
The distribution patterns of deep infaunal species (deep microhabitat by the Pearson correlation coefficient, with values between 0.3 and 0.8,
taxa), shallow to intermediate infaunal and shallow infaunal grouped and the Kaiser-Meyer-Olkin test (KMO) values, which were higher than
together (intermediate microhabitat taxa), and epifaunal to shallow 0.5. All the selected benthic taxa were used to identify the PCs. The
infaunal and epifaunal grouped together (superficial microhabitat taxa), Bartlett sphericity test (0.001) allowed us to reject the null hypothesis
are reported in Figs. 4e-g. (intercorrelation matrix comes from a population in which the variables
From 16.12 to 16.01 Ma superficial microhabitat taxa are dominant are non-collinear) and the KMO of 0.7 indicated that the extracted PCs
with mean percentage values around 45%, followed by intermediate could account for a fair amount of variance. The first two PCs explained
microhabitat taxa with mean percentage values around 25%, and by about 39.2% of the variance, the magnitude of loading factors indicating
deep microhabitat taxa which are clearly subordinate with mean per­ the degree of correlation that links the single variables are reported in
centage values around 8%. From 16.01 to 15.91 Ma, just below the base Table 3.
of the transitional thick dark-grey interval, superficial microhabitat taxa The first Principal component (PCA1) explains 25.61% of the total
percentage values abruptly fall from 47% to 12% (the lowest value is variance. The highest positive loading characterizes the C. ungerianus
10.19% at 15.96 Ma), showing mean percentage values around 15%, (0.71), C. cicatricosus (0.69), and C. pachyderma (0.68) species (Table 3),
and increasing only in the uppermost part of the interval, while inter­ which are plotted together in Fig. 5b, followed by H. boueana (Table 3)
mediate microhabitat taxa show an opposite trend increasing up to 64% and U. peregrina (Fig. 5c; Table 3). The negative loading typifies
(maximum value along the section) at 15.94 Ma and then decreasing C. subhaidingerii/H. dutemplei (− 0.63) and S. reticulata (− 0.53). The
again up to the top of the interval; deep microhabitat taxa increases up PCA1 score plot shows a fluctuating trend with values ranging between
to 21% at 15.96 Ma followed by constant values around 10% up to the − 1.6 and 1.9 from the base up to the top of the section (Fig. 5a).
top of the interval. From 15.91 to 15.62 Ma the three microhabitats Nevertheless, some moderate changes can be observed in the intervals
curves alternate increasing and decreasing trends. In particular, super­ from 16.01 to 15.91 Ma and from 15.86 to 15.83 Ma where the PCA1
ficial microhabitat taxa increase and become dominant in the intervals negative scores shift towards higher values (respectively: − 0.7 and −
15.91–15.86 Ma and 15.83–15.71 Ma, where deep and intermediate 0.6) (Fig. 5a).
microhabitat taxa decrease; between 15.86 and 15.83 Ma, and 15.71 The second principal component (PCA2) explains 13.62% of the total
and 15.62 Ma deep and intermediate microhabitat taxa increase, while variance. The taxa C. pachyderma, C. subhaidingerii/H. dutemplei,
superficial microhabitat taxa decrease and become subordinate to in­ C. ungerianus, S. reticulata and S. carinatus, all belonging to Group A
termediate microhabitat taxa. Upwards from 15.62 to 15.37 Ma super­ (Fig. 5e), have positive loading (Table 3), while the species B. striata,
ficial microhabitat taxa increase and become dominant again with mean B. spathulata and U. cylindrica gaudrynoides, all belonging to Group B,
values around 50%, while intermediate and deep microhabitats taxa have negative loading (Table 3). The PCA2 score plot shows four main
decrease showing mean values respectively around 25% and 10%. intervals with positive PCA2 score: from 16.12 to 16.01 Ma, from 15.91
Finally, from 15.37 Ma to the top of the section superficial microhabitat to 15.86 Ma, from 15.82 to 15.71 Ma and from 15.62 to 15.37 Ma,

7
B. Russo et al.
8

Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722


Fig. 3. - Vertical distribution patterns of the benthic foraminiferal taxa occurring with an abundance >4% along the St. Peter’s Pool section; for each taxon the vertical distribution pattern is plotted together with the 3
point moving average. On the left of the lithologic column the planktonic foraminifera and calcareous nannoplancton bioevents are reported. X-axis scaling is variable to better highlight the changes in abundance of the
different taxa along the section. The grey stripe corresponds to the thick dark-grey interval occurring between 10.2 and 11.9 m.
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

Fig. 4. From left to right curves plotted in time domain (Ma) of a: Group A (oxic/oligotrophic taxa + oxic/oligo-mesotrophic opportunisting behaving species); b:
Group B (hypoxic/eutrophic species); c: oxic/oligotrophic taxa; d: oxic/oligo-mesotrophic opportunisting behaving species; e: deep microhabitat taxa (deep infaunal
species); f: intermediate microhabitat taxa (shallow to intermediate infaunal + shallow infaunal taxa); g: superficial microhabitat taxa (epifaunal to shallow infaunal
+ epifaunal taxa); h: curve of relative oxygen level at the sea floor (O2 index) (Kouwenhoven and van der Zwaan, 2006); i: %P (van Hinsbergen et al., 2005); j-k:
oxygen (δ18O) and carbon (δ13C) stable isotopes of the benthic foraminifera H. dutemplei. For each plot the original data curve is reported together with the 3 point
moving average. The thin black lines represent the correlation between meters (lithologic column) and time (Ma). On the left of the lithologic column the planktonic
foraminifera and calcareous nannoplancton bioevents are reported. For the list of taxa belonging to Group A, Group B, and to the different microhabitats see Table 2.
The grey stripe corresponds to the thick dark-grey interval occurring between 10.2 and 11.9 m.

16.01 Ma O2 index has mean values around 1.5, moving between a


Table 3
minimum value of 1.19 (16.10 Ma) and a maximum of 1.88 (16.09 and
Factor loadings of first (PCA1) and second (PCA2) Principal Components.
16.08 Ma). From 16.01 to 15.91 Ma, just below the base of the transi­
Species PCA1 PCA2 tional thick dark-grey interval, O2 index values abruptly fall from 1.65 to
Bolivina reticulata − 0.32 − 0.19 1.00 (lowest value at 16.00 Ma), showing mean values around 1.08, and
Bolivina spathulata − 0.29 − 0.43 increasing only at the uppermost part of the interval. From 15.91 to
Bulimina striata 0.04 − 0.60 15.62 Ma O2 index values alternate increasing (15.91 to 15.86 Ma; 15.83
Cibicides pachyderma 0.68 0.44
Cibicidoides cicatricosus 0.69 0.25
to 15.71 Ma) and decreasing (15.86 to 15.83 Ma; 15.71 to 15.62 Ma).
Cibicidoides subhaidingerii/ Heterolepa dutemplei − 0.63 0.46 Upwards from 15.62 to 15.37 Ma O2 index values increase with mean
Cibicidoides ungerianus 0.71 0.57 values around 1.7. Finally from 15.37 Ma to the top of the section O2
Gyroidina spp. 0.51 − 0.26 index slightly decreases to mean values of 1.55.
Hanzawaia boueana 0.63 − 0.20
Lenticulina spp. − 0.58 0.08
Melonis affinis − 0.44 − 0.02 4.5. Plankton/benthos ratio (%P)
Oridorsalis umbonatus 0.35 0.28
Pullenia bulloides 0.53 − 0.05
Plankton/benthos ratio (%P) was calculated for all the samples, since
Siphonina reticulata − 0.53 0.34
Spirorutilus carinatus − 0.37 0.41 there was no evidence of downslope transport or reworking. As shown in
Uvigerina cylindrica gaudrynoides 0.14 − 0.64 Fig. 4i, %P values prevalently range from 74 to 92%, with lower values
Uvigerina peregrina 0.55 − 0.21 occurring respectively at 15.36 Ma (66%), 15.37 Ma (52%), 16.01 Ma
(63%), 16.09 Ma (67%) and 16.12 Ma (48%).
interbedded by four intervals, characterized by negative PCA2 scores:
from 16.01 to 15.91 Ma, from 15.86 to 15.82 Ma, from 15.71 to 15.62 4.6. Stable isotopes
Ma and from 15.37 to 15.36 Ma (Fig. 5d).
The benthic δ18OH. dutemplei and δ13CH. dutemplei isotope records are
reported in Fig. 4j and Fig. 4k, respectively. The δ18OH. dutemplei signal
4.4. Paleoxygenation index (O2 index) ranges between − 2.5 and 0.4‰. From the base of the section to 16.07
Ma, the δ18O curve does not show any noteworthy change, while from
The taxa used to calculate O2 index and their arrangement in three 16.07 to 15.96 Ma, it shows a trend towards heavier values from − 1.7‰
groups, each considered to be related to the increasing content of oxygen to about 0.5‰. Upwards to 15.64 Ma, δ18OH. dutemplei signal documents a
levels at sea floor, are reported in Table 4. cyclic trend with lighter (from 15.96 to 15.86 Ma and from 15.83 to
O2 index values fluctuate along the section (Fig. 4h). From 16.12 to 15.77 Ma) and heavier values (from 15.86 to 15.83 Ma and from 15.77

9
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

this basal interval values increase, showing a peak at 16.01 Ma (2.2‰);


then from 16.01 to 15.77 Ma values decrease but they maintain slightly
higher than in the basal interval. From 15.77 to 15.47 Ma δ13CH. dutemplei
values decrease reaching the minimum at 15.49 Ma. In the topmost part
of the section values show low-amplitude fluctuations with no major
shifts observed.

5. Discussion

5.1. Deep water environmental changes

The S. Peter’s Pool section is characterized by fluctuations in abun­


dance of the superficial microhabitat oxic taxa (C. subhaidingerii/H.
dutemplei and S. reticulata), typical of oligotrophic bottom conditions, of
the deep infaunal hypoxic species B. spathulata, and of the intermediate
microhabitat hypoxic species (U. peregrina and subordinately B. striata
and U. cylindrica gaudrynoides), typical of eutrophic bottom conditions.
In addiction the occurrence of the superficial microhabitat oxic oppor­
tunistic behaving species (C. pachyderma, C. ungerianus and
C. cicatricosus), typical of oligo-mesothrophic bottom conditions, show a
significant trend. These latter three species generally show their highest
abundances when the other oxic and hypoxic species have similar total
relative abundances, testifying an increasing of organic matter flux to
Fig. 5. a: sample score pattern on the first principal component (PCA1); ver­ the sea-floor in still oxygenated bottom waters masses. In these envi­
tical distribution patterns of b: oxic/oligo-mesotrophic opportunistic behaving ronmental bottom conditions the deep infaunal species B. spathulata is
species (C. ungerianus + C. cicatricosus + C. pachyderma), c: U. peregrina; d: the only hypoxic eutrophic species which does not increase. We believe
sample score pattern on the second principal component (PCA2); e: vertical that the increased organic matter flux to the sea-floor, since in still
distribution patterns of Group A (see Table 1). For each plot the original data oxygenated bottom waters, characterizes a microhabitat where the oxic/
curve is reported together with the 3 point moving average. On the left of the
oligo-mesotrophic opportunistic behaving species are well adapted. The
lithologic column the planktonic foraminifera and calcareous nannoplancton
increasing in abundance of these species in such environmental bottom
bioevents are reported. The grey stripe corresponds to the thick dark-grey in­
terval occuring between 10.2 and 11.9 m. conditions testifies their opportunistic behavior, in agreement with
Verhallen (1991) and Frezza et al. (2005), who indicate that opportu­
nistic species increase with the increasing of the organic matter content
Table 4
and the decreasing of oxygen content at the sea floor. The bibliographic
Groups of benthic taxa arranged according to their relative increasing sensibility
to the oxygen content at the sea floor, used to construct the oxygen index references relative to ecological/palaecological meaning of the species
(Kouwenhoven and van der Zwaan, 2006). The three groups correspond to three are reported in section 3.2, Table 1.
increasing oxygen levels: low (Group 1), intermediate (Group 2), and high Distribution patterns of benthic species together with isotopic data
(Group 3) level of oxygen bottom waters. and statistical analysis results allowed us to identify five main intervals
Group 1 Group 2 Group 3
from SPP1 to SPP5 (Fig. 6) showing changes in the structure of the
benthic foraminiferal assemblage, their microhabitats and the evolution
Bolivina reticulata Cibicides pachyderma Anomalinoides helicinus
of bottom water conditions. The palaeoecological changes observed
Bolivina spathulata Cibicidoies cicatricosus Cibicidoides
Bulimina striata Cibicidoides ungerianus subhaidingerii/Heterolepa have also been related to the sea-level variations from the base to the top
dutemplei of the section (16.12–15.36 Ma) reported by Miller et al. (2020).
Cassidulina laevigata Globocassidulina Siphonina reticulata Between 16.12 and 16.01 Ma (interval SPP1) the general dominance
subglobosa
of Group A (preferring superficial microhabitat) (Fig. 6a), together with
Hanzawaia boueana Gyroidina spp. Spirorutilus carinatus
the relatively higher O2 index values (Fig. 6e), the meaningful presence
Uvigerina cylindrica Lenticulina spp. of Group B (Fig. 6a), the subordinate occurrence of the deep micro­
gaudrynoides habitat species B. spathulata (showing the lowest percentage values
Uvigerina peregrina Oridorsalis umbonatus starting from 16.08 Ma) (Fig. 6c) and the δ13CH. dutemplei values which
Pullenia bulloides
start to increase from 16.07 Ma, showing a peak at 16.01 Ma (Fig. 6g),
Data from: Alavi, 1988; Corliss, 1991; Kaiho, 1994, 1999; Schmiedl et al., 1997; suggest relatively oxygenated bottom water conditions with moderate
Fontanier et al., 2002; Spezzaferri et al., 2002, 2013; Murray, 2003, 2006; Rögl organic matter fluxes. This interpretation is also supported by 1) the
and Spezzaferri, 2003; Schmiedl et al., 2003; De Man, 2006; Kouwenhoven and fluctuation of PCA1 (Fig. 5a), which can be considered as the surface
van der Zwaan, 2006; Jorissen et al., 2007; Martins et al., 2007; Oblak, 2007; productivity trend, since it well correlate with the distribution patterns
Russo et al., 2007; Abu-Zied et al., 2008; Drinia, 2009; Melki et al., 2009; Alve,
of U. peregrina and B. spathulata (Fig. 3), and by 2) PCA2 curve (Fig. 6d),
2010; Ied et al., 2011; Grunert et al., 2012; Rai and Singh, 2012; Sgarrella et al.,
which can be interpreted as the oxygen content of bottom water trend,
2012; Singh et al., 2015; Patarroyo-Camargo and Martinez-Rodriguez, 2013;
Székely et al., 2017; Drinia, 2018; Milker et al., 2019; Pascual et al., 2020; Pérez-
since it mirrors the patterns of C. subhaidingerii/H. dutemplei, S. reticulata
Asensio et al., 2020. The ecological meaning of the different taxa is reported in (Fig. 3) and 3) O2 index (Fig. 6e). Sporadic episodes of higher oxygen
the section 3.2, Table 1. content are testified by peaks of C. subhaidingerii/H. dutemplei (Fig. 3),
and the superficial microhabitat taxa (Fig. 6c), and by relatively higher
values of O2 index (Fig. 6e). On the contrary, peaks in abundance of the
to 15.64 Ma). A gradual decreasing towards lighter values (− 2‰) is intermediate microhabitat taxa (B. striata, C. laevigata, U. cylindrica
observed from 15.64 to 15.52 Ma, followed by a trend vs. heavy values gaudrynoides, U. peregrina, Fig. 3) document short periods of enhanced
up to the top of the section. productivity in still oxygenated bottom waters, as also highlighted by
The δ13CH. dutemplei values range between 0.3 and 2.2‰. In particular δ13CH. dutemplei signal (Fig. 6g). In SPP1 interval M. affinis and C. laevigata
the curve does not show variations from the base to 16.07 Ma. Above display an increasing trend from 16.08 to 16.06 Ma, showing some of the

10
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

Fig. 6. From left to right: a) total relative abundance curves of Group A (red) and Group B (blue); b) total relative abundance curves of oxic/oligo-mesotrophic
opportunistic behaving species (green), oxic/oligotrophic taxa (red), and hypoxic/eutrophic species (blue); c) total relative abundance curves of deep microhab­
itat taxa (black), intermediate microhabitat taxa (blue), and superficial microhabitat taxa (red); d) sample score pattern on the second principal component (PCA 2);
e) curve of relative oxygen level at the sea floor (O2 index) (Kouwenhoven and van der Zwaan, 2006); f) vertical distribution pattern of Melonis affinis; g)-h) oxygen
(δ18O) and carbon (δ13C) stable isotopes curves of the benthic foraminifera Heterolepa dutemplei; i) sea level oscillations curve (Miller et al., 2020); intervals (SPP1-
SPP5). On the left of the lithologic column, the planktonic foraminifera and calcareous nannoplancton bioevents are reported. For the list of taxa belonging to A and B
groups, oxic/oligo-mesotrophic opportunistic behaving species, oxic/oligotrophic taxa, and hypoxic/eutrophic species, deep microhabitat taxa, intermediate
microhabitat taxa, and superficial microhabitat taxa see Table 2. The grey stripe corresponds to the thick dark-grey interval occurring between 10.2 and 11.9 m. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

highest percentage values recorded along the section (Fig. 6f). This (interval SPP1) to low oxygen bottom water masses with a clear increase
increasing trend correlates with δ18OH. dutemplei heavier values starting of the productivity. These new conditions at the sea floor are also
from 16.07 Ma (Fig. 6h), testifying the beginning of a relative cooling of associated with a cooling phase testified by a marked shift vs. heavy
bottom water masses. values in δ18OH. dutemplei signature (Fig. 6h) and by the increasing trend of
At 16.01 Ma, the base of the thick dark-grey interval marks a drastic M. affinis and C. laevigata (Fig. 6f). This cooling phase should correspond
change in the benthic foraminiferal assemblage, because Group A is to the well-known glaciation event Mi2 (Miller et al., 1991), reflecting
replaced by Group B (Fig. 6a); the intermediate microhabitat taxa substantial ice volume increases near the bases of Zones Mi2 (at about
become dominant (Fig. 6c) and the deep microhabitat taxa increase 16.1 Ma), and correlates with the sea level falling (Fig. 6i, Miller et al.,
(Fig. 6c), while the superficial microhabitat taxa abruptly decrease 2020). The sea level falling trend, as well as δ18OH. dutemplei (Fig. 6h) and
(Fig. 6c) with percentages values generally similar or lower (in corre­ M. affinis trends (Fig. 6f), matches also the decreasing trend of Group A,
spondence of the thick dark-grey interval) than the deep microhabitat the opposite increasing trend of Group B (Fig. 6a), and the increasing
taxa (Fig. 6c). U. cylindrica gaudrynoides, species indicative of high pri­ trend of δ13CH. dutemplei (Fig. 6g). These conditions, together with the
mary productivity (e.g. Székely et al., 2017) and proliferating in stressed significant presence of U. cylindrica gaudrynoides and C. laevigata, could
and nutrient-rich environments (Drinia, 2018), characterizes this in­ be due to an increase of the primary productivity, caused by the inten­
terval. Cassidulina laevigata, a high stress, dysoxic and high organic sification of the runoff determined by the sea level fall related to the
matter tolerant species (e.g. Spezzaferri et al., 2013), evidencing the cooling phase.
occurrence of cool waters with low-oxygen and high organic matter Between 15.91 and 15.62 Ma (interval SPP3) the fluctuations of the
content (Pascual et al., 2020), displays a significant increase with some superficial microhabitat Group A, and intermediate/deep microhabitat
of its highest percentage values. The increase of Group B and the trends Group B, which become either dominant or subordinate (Fig. 6a), imply
of C. laevigata and U. cylindrica gaudrynoides may be the result of an a transition from a more stable and eutrophic conditions (SPP2) to un­
increased organic flux at the seafloor (e.g. Jorissen et al., 1995). In stable oligo-mesotrophic/eutrophic conditions (SPP3). This interpreta­
addition, the PCA1 negative scores shift towards higher values (Fig. 5a), tion is supported by the increase in abundance of the superficial
while O2 index abruptly decreases and reaches its minimum values microhabitat, oxic/oligo-mesotrophic opportunistic behaving species
(Fig. 6e). PCA2 becomes negative (Fig. 6d), whereas δ13CH. dutemplei (C. pachyderma, C. cicatricosus and C. ungerianus), species more
shows a peak followed by a decreasing trend but with values slightly competitive in unstable environmental conditions, from 15.91 to 15.71
higher than in the SPP1 interval (Fig. 6g). All these signals evidence, Ma (Fig. 6b). These changed bottom water conditions are also docu­
between 16.01 and 15.91 Ma (interval SPP2), the transition from a quite mented by the trend in O2 index (Fig. 6e), PCA2 (Fig. 6d) and partially
oxygenated with moderate organic matter fluxes to the sea floor by PCA1 curves (Fig. 5a), as well as by the trend of δ18OH. dutemplei signal

11
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

(Fig. 6h), suggesting relatively more oxygenated and warmer (from 500 and 1000 m. The co-occurrence of the depth markers C. cicatricosus
15.91 to 15.86 Ma, and from 15.83 to 15.77 Ma) or more productive, less together with B. reticulata, H. dutemplei and H. boueana confirm and
oxygenated and cooler (from 15.86 to 15.83 Ma and from 15.77 to refine the depth interval calculated with the %P. Cibicidoides cicatricosus
15.62 Ma) bottom water masses. In SPP3 interval M. affinis shows an has its upper and lower bathymetric limits at 600 m and 1000 m
increasing trend from 15.78 to 15.62 Ma (Fig. 6f) correlating to the respectively (Katz and Miller, 1993); while the other species have their
δ18OH. dutemplei vs. heavier values (Fig. 6h). In SPP3 interval, the varia­ lower limit at 1000 m (Kouwenhoven, 2000; Russo et al., 2002; Drinia
tions of the sea level correlate with changes in Group A, M. affinis and et al., 2004), then a middle bathyal paleodepth (van Morkhoven et al.,
δ18OH. dutemplei only in the upper part (from 15.71 to 15.63 Ma), where 1986), ranging from 600 m to 1000 m is estimated.
the oxic/oligo-mesotrophic opportunistic behaving species are less
abundant than the interval 15.91–15.71 Ma (Fig. 6). 6. Conclusions
Between 15.62 and 15.37 Ma (interval SPP4) the palaecological
bottom conditions are similar to the SPP1 interval, since the superficial The integrated study on benthic foraminiferal assemblages and Ox­
microhabitat Group A (Fig. 6a) dominates again, and the oxic/oligo- ygen and Carbon stable isotopes were carried out on the upper Burdi­
mesotrophic opportunistic behaving species occur even if with a galian - lower Langhian S. Peter’s Pool section (Malta Island, Central
discontinuous trend (Fig. 6b), Group B slightly decreases, suggesting Mediterranean Sea).
relatively stable oxic conditions with moderate organic matter fluxes. An The micropaleontological and geochemical data collected from the
increasing trend is recorded by O2 index curve (Fig. 6e), PCA2 becomes section allow us to hypothesize the following palaeoecological changes
positive (Fig. 6d), while the intermediate and deep microhabitat Group in the bottom water masses:
B decreases (Fig. 6a) as well as δ13CH. dutemplei signal (Fig. 6g). These
changes suggest a partial restoring of the SPP1 palaeoecological bottom - from 16.12 to 16.01 Ma (interval SPP1): relatively oxygenated
water conditions. δ18OH. dutemplei trend initially decreases vs. lighter bottom conditions with moderate organic matter fluxes;
values and gradually increases towards heavier values after 15.52 Ma - from 16.01 (base of the thick dark-grey lithological interval) to
(Fig. 6h), in good agreement with the decreasing followed by the 15.91 Ma (interval SPP2): transition to cooler bottom water masses
increasing trends of C. laevigata (initially decreasing and then increasing with lower oxygen content and a clear increase of the productivity. In
from 15.54 Ma, Fig. 3) and M. affinis (initially decreasing and then this interval is recorded the major benthic foraminiferal turnover,
increasing from 15.50 Ma, Fig. 6f). These trends also correlate with the that is well recognized also in the isotope records. These conditions
initial rising followed by the sea level fall (Fig. 6i). It is worth to point are associated with a sea level fall, well correlated with a cooling
out that δ18OH. dutemplei signal correlates with the Group A and Group B phase, documented by δ18OH. dutemplei signal, M. affinis trend and the
trends only up to 15.52 Ma (lighter δ18OH. dutemplei values, Group A trends of the other main palaeoecological bottom water indicators.
increasing trend and Group B decreasing trend), but upwards to 15.37 We believe that these conditions are related to an increase of the
Ma, coherently with the shift of δ18OH. dutemplei signal towards heavier primary productivity, caused by an intensification of the runoff
values, Group A and Group B do not invert their trends. We have not determined by the falling of the sea level, following the cooling;
found an explanation for this mismatch. - from 15.91 to 15.62 Ma (interval SPP3): succeeding of relatively
In the uppermost part of the section, from 15.37 to 15.36 Ma (in­ more oxygenated and warmer bottom water masses (from 15.91 to
terval SPP5): 1) the slight decrease in abundance of the superficial 15.86 Ma and from 15.83 to 15.71 Ma) with more productive, less
microhabitat Group A (Fig. 6a) - mainly due the drastic reduction of the oxygenated and cooler water masses (from 15.86 to 15.83 Ma and
oxic/oligo-mesothrophic opportunistic behaving species (Fig. 6b), 2) the from 15.71 to 15.62 Ma);
moderate increase of the intermediate/deep microhabitat Group B - from 15.62 to 15.37 Ma (interval SPP4): restoring of the initial
(Fig. 6a) - essentially due to the increase of B. spathulata (Fig. 6c), 3) the bottom conditions recorded in the lower part (interval SPP1) of the
shift of PCA2 towards negative values (Fig. 6d), 4) the slight decrease of section, characterized by relatively oxygenated bottom water masses
O2 index (Fig. 6e), suggest a little increase of the organic matter content with moderate organic matter fluxes. These conditions are associated
in relatively less oxygenated bottom water masses with respects to SPP1 with an initial rising, followed by a falling of the sea level, well
and SPP4 intervals. correlated with δ18OH. dutemplei signal initially lighter and heavier
Summarizing, the benthic foraminiferal changes reflect different upwards;
deep water environmental conditions between 16.12 and 15.36 Ma (late - from 15.37 to 15.36 Ma (interval SPP5): little increase of the organic
Burdigalian to early Langhian) corresponding to warm climate condition matter content in relatively less oxygenated bottom water masses
linked to the onset of the MCO (e.g. Holbourn et al., 2015; Prista et al., with respects to SPP1 and SPP4 intervals.
2015).
Thomas (2007) reported that middle Miocene benthic foraminiferal Plankton/benthos ratio (%P) and the presence of some benthic spe­
faunal turnover, occurred during episodes of global cooling, may reflect cies with a well-defined depth distribution range allowed us to estimate
the increased seasonality of primary productivity and increased delivery a middle-bathyal paleodepth, ranging from 600 m to 1000 m during
of labile organic matter to the sea floor. According to this interpretation, MCO (between 17 and 15 Ma).
we suggest that in the St. Peter’s Pool section the fluctuations of “oxic/ It is worth to point out that the major benthic foraminiferal turnover
oligotrophic”, “hypoxic/eutrophic” and “oxic/oligo-mesotrophic occurred between 16.01 (base of the thick dark-grey lithological inter­
opportunistic behaving” taxa are indicative of increased seasonality, val) and 15.91 Ma can be related to a cooling phase corresponding to the
which in turn affect food flux and oxygenation at the sea-floor. onset of Miller’s event Mi2 (Miller et al., 1991, 2020).
Concluding, the main variations detected in the benthic foraminif­
5.2. Paleobathymetry eral assemblages suggest different paleoenvironmental conditions
characterized either by relatively oxygenated, warmer bottom water
Foresi et al. (2011) estimated an upper-bathyal paleodepth for St. masses and moderate organic matter fluxes, or less oxygenated, cooler
Peter’s Pool section deposition, based on a preliminary study of benthic bottom water masses with an increase in productivity, correlating with
foraminifera. More detailed analyses reported in this work allowed us to Miocene Climatic Optimum (MCO). We suggest that these paleoenvir­
estimate the paleodepth with a greater accuracy using the %P (van onmental changes observed in the St. Peter’s Pool section are indicative
Hinsbergen et al., 2005) and marker species with well-defined depth of increased seasonality, which in turn affect food flux and oxygenation
distribution range. The values of the %P curve, mainly ranging from at the sea-floor.
74% to 93% (Fig. 3), permit to estimate a paleodepth ranging between

12
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

Declaration of Competing Interest dell’Università e della Ricerca (PRIN Project 2006, Scientific coordi­
nator Silvia Iaccarino) and by Regional Committee on Mediterranean
The authors declare that they have no known competing financial Neogene Stratigraphy (RCMNS) fund 2009 of the Consiglio Nazionale
interests or personal relationships that could have appeared to influence delle Ricerche (CNR) of Italy.
the work reported in this paper. We thank Kenneth Miller (Rutgers University) for the sea level data.

Acknowledgments

This work was financially supported by Ministero dell’Istruzione,

Appendix I

List of the examined samples: sample, stratigraphic position (distance in meters from the base of the section), sample spacing (distance in cen­
timeters between one sample and the next), lithology (DBHL = dark beige hardened limestone; DGML = dark grey marly limestone; DGCM = dark grey
calcareous marl; WML = whitish marly limestone; WCL = whitish calcareous limestone; DBML = dark beige marly limestone; DBCM = dark beige
calcareous marl; LBML = light beige marly limestone; LBCM = light beige calcareous marl).

Sample Stratigraphic position (m) Sample spacing (cm) Lithology Sample Stratigraphic position (m) Sample spacing (cm) Lithology

267bis 30.99 54 DBHL 126 16.13 31 DBML


264 30.45 8 WML 124 15.82 27 DBML
263 30.37 11 WML 122 15.55 9 DBML
261 30.26 23 WCL 121 15.46 25 DBHL
258 30.03 37 WCL 119 15.21 19 DBML
254bis 29.66 4 WCL 117 15.02 24 DBHL
254 29.62 38 WCL 115 14.78 23 DBML
251 29.24 29 WML 113 14.55 14 DBHL
248 28.95 30 WCL 328 14.41 66 DBML
245 28.65 6 WCL 107 13.75 9 DBHL
244 28.59 8 WCL 106 13.66 21 DBHL
242 28.51 46 WCL 104 13.45 11 DBHL
237 28.05 28 DBCM 103 13.34 9 DBML
234 27.77 41 DBCM 102 13.25 10 DBHL
231 27.36 11 DBCM 101 13.15 23 DBML
230 27.25 29 DBCM 99 12.92 13 DBHL
227 26.96 32 DBCM 98 12.79 24 DBML
224 26.64 37 DBHL 95 12.55 14 DBHL
220 26.27 8 DBCM 94 12.41 11 DBHL
219 26.19 18 DGML 93bis 12.3 23 DBHL
217 26.01 13 DBCM 91 12.07 28 DBML
216 25.88 23 DBCM 89 11.79 75 DGCM
214 25.65 31 DBCM 84 11.04 14 WML
211 25.34 14 DBCM 281 10.90 30 WML
209 25.20 27 DBHL 81 10.60 18 DGCM
207 24.93 20 DBCM 80 10.42 16 DGCM
205 24.73 33 DBCM 78 10.26 29 DGCM
203 24.40 12 DBCM 76 9.97 28 LBCM
202 24.28 39 DBCM 74 9.69 20 LBCM
199 23.89 15 DBHL 72 9.49 27 LBCM
197 23.74 16 DBHL 70 9.22 20 LBCM
196 23.58 33 DBCM 68 9.02 10 DBHL
193 23.25 26 DBCM 67 8.92 12 LBML
191 22.99 8 DBCM 66 8.80 22 LBML
190 22.91 21 DBCM 64 8.58 41 LBML
188 22.70 15 DBCM 61 8.17 35 LBCM
187 22.55 23 DBCM 59 7.82 19 LBML
185 22.32 20 DBCM 57 7.63 31 LBML
183 22.12 28 DBCM 55 7.32 7 LBCM
181 21.84 32 DBHL 54 7.25 45 LBCM
178 21.52 22 DBCM 50 6.80 18 LBCM
176 21.30 10 DBHL 48ter 6.62 44 DBHL
175 21.20 9 DBCM 342 6.18 31 LBCM
174 21.11 22 DBCM 336 5.87 20 DGML
172 20.89 20 DBCM 332 5.67 10 LBCM
171 20.69 15 DBHL 330 5.57 34 LBCM
169 20.54 8 DBCM 36 5.23 11 LBCM
168 20.46 7 DBCM 35 5.12 29 LBCM
167 20.39 21 DBHL 33bis 4.83 21 LBCM
165 20.18 21 DBCM 32 4.62 9 DBHL
162 19.97 25 DBCM 31 4.53 26 LBCM
159 19.72 14 DBCM 29 4.27 30 LBML
157 19.58 12 DBCM 26 3.97 17 LBCM
(continued on next page)

13
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

(continued )
Sample Stratigraphic position (m) Sample spacing (cm) Lithology Sample Stratigraphic position (m) Sample spacing (cm) Lithology

156 19.46 9 DBCM 25 3.80 26 LBCM


155 19.37 32 DBCM 23 3.54 18 LBML
365 19.05 30 DBHL 22 3.36 40 LBCM
359 18.75 10 DBHL 20 2.96 14 LBCM
357 18.65 54 DBHL 19bis 2.82 22 DBHL
141 18.11 9 DBML 18 2.60 37 LBCM
140 18.02 29 DBML 15 2.23 41 LBML
138 17.73 16 DBML 12 1.82 5 LBCM
137 17.57 21 DGML 11 1.77 40 LBCM
135 17.36 39 DBML 8 1.37 28 DGML
133 16.97 16 DBHL 6 1.09 26 LBCM
132 16.81 10 DBHL 4 0.83 36 LBML
131 16.71 10 DBML 2bis 0.47 39 DBHL
130 16.61 32 DBHL 1 0.08 8 LBCM
127 16.29 16 DBML

Appendix II - List of the classified species.

Anomalinoides helicinus (Costa) = Nonionina helicina Costa, 1857.


Bolivina reticulata Hantken, 1875.
Bolivina spathulata (Williamson) = Textularia variabilis var. spathulata Williamson, 1858.
Bulimina elongata d’Orbigny, 1846.
Bulimina striata d’Orbigny in Guérin-Méneville,1832.
Cancris auricula (Fichtel and Moll) = Nautilus auricula Fichtel and Moll, 1798.
Cancris oblongus (Williamson) = Rotalina oblonga Williamson, 1858.
Cassidulina carinata Silvestri, 1896 = C. laevigata var. carinata Silvestri, 1896.
Cassidulina laevigata d’Orbigny, 1826.
Cibicides pachyderma (Rzehak) = Truncatulina pachyderma Rzehak, 1886.
Cibicidoides cicatricosus (Schwager) = Anomalina cicatricosa Schwager, 1866.
Cibicidoides subhaidingerii (Parr) = Cibicides subhaidingerii Parr, 1950.
Cibicidoides ungerianus (d’Orbigny) = Rotalina ungeriana d’Orbigny, 1846.
Cibicidoides wuellerstorfi (Schwager) = Anomalina wuellerstorfi Schwager, 1866.
Cribroelphidium albiumbilicatum (Weiss, 1954) = Nonion pauciloculum subsp. albiumbilicatum Weiss, 1954.
Epistominella lecalvezi (Lys and Bourdon) = Pseudoparrella lecalvezi Lys and Bourdon, 1958.
Globobulimina affinis (d’Orbigny) = Bulimina affinis d’Orbigny, 1839.
Globocassidulina subglobosa (Brady) = Cassidulina subglobosa Brady, 1881.
Gyroidina spp.
Hanzawaia boueana (d’Orbigny) = Truncatulina boueana d’Orbigny, 1846.
Heterolepa dutemplei (d’Orbigny) = Rotalina dutemplei d’Orbigny, 1846.
Karreriella sp.
Lenticulina spp.
Melonis affinis (Reuss) = Nonionina affinis Reuss, 1851
Neoeponides schreibersii (d’Orbigny) = Rotalina schreibersii d’Orbigny, 1846.
Oridorsalis umbonatus (Reuss) = Rotalina umbonata Reuss, 1851.
Pullenia bulloides (d’Orbigny) = Nonionina bulloides d’Orbigny, 1846.
Pullenia quadriloba Reuss = Pullenia compressiuscula Reuss var. quadriloba Reuss, 1867.
Pullenia quinqueloba (Reuss) = Nonionina quinqueloba Reuss, 1851.
Pullenia salisburyi R.E. and K.C. Stewart, 1930.
Saracenaria italica Defrance, 1824.
Sigmoilopsis celata (Costa) = Spiroloculina celata Costa, 1855.
Siphonina reticulata (Cžjžek) = Rotalina reticulata Cžjžek, 1848.
Sphaeroidina bulloides d’Orbigny in Deshayes, 1828.
Spirorutilus carinatus (d’Orbigny) = Textularia carinata d’Orbigny, 1846.
Spirosigmoilina tenuis (Cžjžek) = Quinqueloculina tenuis Cžjžek, 1848.
Trifarina bradyi Cushman, 1923.
Uvigerina barbatula Macfadyen, 1930.
Uvigerina cylindrica gaudrynoides Lipparini, 1932.
Uvigerina peregrina Cushman, 1923.
Uvigerina rutila Cushman and Todd, 1941.
Uvigerina semiornata d’Orbigny, 1846.
Uvigerina striatissima Perconig, 1955.

14
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.palaeo.2021.110722.

References the Moira Mounds (Porcupine Seabight, SW of Ireland). Swiss J. Geosci. https://doi.
org/10.1007/s00015-018-0317-4.
Finetti, I., 1982. Structure, stratigraphy and evolution of Central Mediterranean. Boll.
Abu-Zied, R.H., Rohling, E.J., Jorissen, F.J., Fontanier, C., Casford, J.S.L., Cooke, S.,
Geofis. Teor. Appl. 24, 247–426.
2008. Benthic foraminiferal response to changes in bottom-water oxygenation and
Föllmi, K.B., Gertsch, B., Renevey, J.P., De Kaenel, E., Stilles, P., 2008. Stratigraphy and
organic carbon flux in the eastern Mediterranean during LGM to Recent times. Mar.
sedimentology of phosphate-rich sediments in Malta and South-eastern Sicily (latest
Micropaleontol. 67, 46–68.
Oligocene to early late Miocene). Sedimentology 55, 1029–1051.
Alavi, S.N., 1988. Late Holocene deep-sea benthic foraminifera from the Sea of Marmara.
Fontanier, C., Jorissen, F.J., Licari, L., Alexandre, A., Anschutz, P., Carbonel, P., 2002.
Mar. Micropaleontol. 13, 213–237.
Live benthic foraminiferal faunas from the Bay of Biscay: faunal density,
Alve, E., 2010. Benthic foraminiferal responses to absence of fresh phytodetritus: a two
composition, and microhabitats. Deep-Sea Res PT I 49, 751–785.
year experiment. Mar. Micropaleontol. 76, 67–75.
Fontanier, C., Jorissen, F.J., Chaillou, G., David, C., Anschutz, P., Lafon, V., 2003.
Baas, J.H., Schönfeld, J., Zahnet, R., 1998. Mid-depth oxygen drawdown during Heinrich
Seasonal and interannual variability of benthic foraminiferal faunas at 550m depth
events: evidence from benthic foraminiferal community structure, trace-fossil
in the Bay of Biscay. Deep-Sea Res PT I 50, 457–494.
tiering, and benthic δ13C at the Portuguese Margin. Mar. Geol. 152, 25–55.
Fontanier, C., Jorissen, F.J., Anschutz, P., Chaillou, G., 2006. Seasonal variability of
Baldassini, N., Di Stefano, A., 2017. Stratigraphic features of the Maltese Archipelago: a
foraminiferal faunas at 1000m depth in the Bay of Biscay. J. Foramin. Res. 36,
synthesis. Nat. Hazards 86, S203–S231.
61–76.
Baldassini, N., Mazzei, R., Foresi, L.M., Riforgiato, F., Salvatorini, G., 2013. Calcareous
Foresi, L.M., Mazzei, R., Salvatorini, G., Donia, F., 2008. Biostratigraphy and
plankton bio-chronostratigraphy of the Maltese Lower Globigerina Limestone
chronostratigraphy of the Maltese Lower Globigerina Limestone Member
Member. Acta Geol. Pol. 63 (1), 105–135.
(Globigerina Limestone Formation): new preliminary data based on calcareous
Barmawidjaja, D.M., Jorissen, F.J., Puscaric, S., van der Zwaan, G.J., 1992. Microhabitat
plankton. B. Soc. Paleontol. Ital. 46 (2–3), 175–181.
selection by benthic foraminifera in the northeast Adriatic Sea. J. Foramin. Res. 22,
Foresi, L.M., Verducci, M., Baldassini, N., Lirer, F., Mazzei, R., Salvatorini, G., Ferraro, L.,
297–317.
Da Pratom, S., 2011. Integrated stratigraphy of St. Peter’s Pool section (Malta): new
Bellanca, A., Sgarrella, F., Neri, R., Russo, B., Sprovieri, M., Bonaduce, G., Rocca, D.,
age for the Upper Globigerina Limestone Member and progress towards the Langhian
2002. Evolution of the Mediterranean Basin during the late Langhian-early
GSSP. Stratigraphy 8, 125–143.
Serravallian: an integrated paleoceanographic approach. Riv. Ital. Paleontol. S. 108
Foresi, L.M., Baldassini, N., Sagnotti, L., Lirer, F., Di Stefano, A., Caricchi, C.,
(2), 223–239.
Verducci, M., Salvatorini, G., Mazzei, R., 2014. Integrated stratigraphy of the St.
Bernhard, J.M., Sen Gupta, B.K., 1999. Foraminifera of oxygen-depleted environments.
Thomas section (Malta Island): a reference section for the lower Burdigalian of the
In: Sen Gupta, B.K. (Ed.), Modern foraminifera. Kluwer Academic Publishers,
Mediterranean Region. Mar. Micropaleontol. 111, 66–89.
pp. 201–216.
Frezza, V., Bergamin, L., Di Bella, L., 2005. Opportunistic benthic foraminifera as
Bonaduce, G., Barra, D., 2002. The ostracods in the palaeoenvironmental interpretation
indicators of eutrophicated environments. Actualistic study and comparison with the
of the late Langhian - early Serravallian section of Ras il-pellegrin (Malta). Riv. It.
Santernian middle Tiber Valley (Central Italy). Boll. Soc. Paleontol. It. 44 (3),
Paleontol. St. 108 (2), 211–222.
193–201.
Brandano, M., Cornacchia, I., Raffi, I., Tomassetti, L., Agostini, S., 2021. The Monterey
Giannelli, L., Salvatorini, G., 1972. I Foraminiferi planctonici dei sedimenti terziari
Event within the Central Mediterranean area: the shallow-water record.
dell’Arcipelago Maltese. Biostratigrafia del “Globigerina Limestone” I. Atti Soc. Tosc.
Sedimentology 64 (1), 286–310.
Sci. Nat., Mem. Ser A 79, 49–74.
Caulle, C., Koho, K.A., Mojtahid, M., Reichart, G.J., Jorissen, F.J., 2013. Live
Giannelli, L., Salvatorini, G., 1975. I foraminiferi planctonici dei sedimenti terziari
foraminiferal faunas (Rose Bengal stained) from the northern Arabian Sea: links with
dell’arcipelago Maltese. I Biostratigrafia di “Blue-clay”, “Green-sands” e “Upper
bottom-water oxygenation. Biogeosci. Discuss. 10, 15257–15304.
Globigerina Limestone”. Atti Soc. Tosc. Sci. Nat., Mem. Ser A 82, 1–24.
Caulle, C., Mojtahid, M., Gooday, A.J., Jorissen, F.J., Kitazato, H., 2015. Living (Rose
Gooday, A.J., 1988. A response by benthic foraminifera to the deposition of
Bengal stained) benthic foraminiferal faunas along a strong bottom-water oxygen
phytodetritus in the deep sea. Nature 332 (6159), 70–73.
gradient on the Indian margin (Arabian Sea). Biogeosci. Discuss. 12, 3245–3282.
Gooday, A.J., 1993. Deep-sea benthic foraminiferal species which exploit phytodetritus:
Corliss, B.H., 1985. Microhabitats of benthic foraminifera within deep-sea sediments.
characteristic features and controls on distribution. Mar. Micropaleontol. 22,
Nature 314, 435–438.
187–205.
Corliss, B.H., 1991. Morphology and microhabitat preferences of benthic foraminifera
Gooday, A.J., 1994. The biology of deep-sea foraminifera: a review of some advances and
from the northwest Atlantic Ocean. Mar. Micropaleontol. 17, 195–236.
their applications in paleoceanography. Palaios 9 (1), 14–31. Published by: SEPM
Davis, J.C., 1973. Statistics and Data Analysis in Geology. Incorporation, New York, John
Society for Sedimentary Geology.
Wiley & ons.
Grunert, P., Soliman, A., Ćorić, S., Roetzel, R., Harzhauser, M., Piller, W.E., 2012. Facies
De La Vara, A., Meijer, P., 2016. Response of Mediterranean circulation to Miocene
development along the tide-influenced shelf of the Burdigalian Seaway: an example
shoaling and closure of the Indian Gateway: a model study. Palaeogeogr. Palaeoclim.
from the Ottnangian stratotype (early Miocene, middle Burdigalian). Mar.
442, 96–109.
Micropaleontol. 84-85, 14–36.
De Man, E., 2006. Benthic foraminifera biofacies analysis and stable isotopes of the
Grunert, P., Rosenthal, Y., Jorissen, F., Holbourn, A., Zhou, X., Piller, W.E., 2018. Mg/Ca-
Middle Eocene to Oligocene successions in the southern North Sea Basin. In: Tools
temperature calibration for costate Bulimina species (B. costata, B. inflata,
for stratigraphy and for reconstruction of extreme climate changes. University of
B. mexicana): a paleothermometer for hypoxic environments. Geochim. Cosmochim.
Leuven, Belgium. Unpublished Ph.D. Thesis. 375 pp.
Acta 220, 36–54. https://doi.org/10.1016/j.gca.2017.09.021.
De Rijk, S., Jorissene, F.J., Rohling, E.J., Troelstra, S.R., 2000. Organic fux control on
Grunert, P., Skinner, L., Hodell, D., Piller, W.E., 2015. A micropalaeontological
bathymetric zonation of Mediterranean benthic foraminifera. Mar. Micropaleontol.
perspective on export productivity, oxygenation and temperature in NE Atlantic
40, 151–166.
deep-waters across Terminations I and II. Glob. Planet. Chang. 131, 174–191.
Di Bella, L., Frezza, V., Conte, A.M., Chiocci, F.L., 2015. Benthic foraminiferal
Gruszczynski, M., Marshall, J.D., Goldring, R., Coleman, M.L., Malkowski, L.,
assemblages in active volcanic area of the Azores Islands (North Atlantic Ocean).
Gazdzicka, E., Semil, J., Gatt, P., 2008. Hiatal surfaces from the Miocene Globigerina
Ital. J. Geosci. 134 (1), 50–59. https://doi.org/10.3301/IJG.2014.22.
Limestone Formation of Malta: biostratigraphy, sedimentology, trace fossils and
Drinia, H., 2009. Palaeoenvironmental reconstruction of the Oligocene Afales Basin,
early diagenesis. Palaeogeogr. Palaeoclim. 270 (2–3), 239–251.
Ithaki island, western Greece. Cent. Eur. J. Geosci. 1 (1), 1–18. https://doi.org/
Guerreiro, C., Rosa, F., Oliveira, A., Cachão, M., Fatela, F., Rodrigues, A., 2009.
10.2478/v10085-009-0001-z.
Calcareous nannoplankton and benthic foraminiferal assemblages from the Nazaré
Drinia, H., 2018. Palaeoenvironmental significance of a late Miocene benthic
Canyon (Portuguese continental margin): preliminary results. In: From Deep-sea to
foraminifera fauna from Apostoli Formation, Central West Crete, Greece. Bull. Geol.
Coasetal Zones: Methods and Techniques for Studying Paleoenvironments IOP
Soc. Greece 34 (2), 627–634. https://doi.org/10.12681/bgsg.17113.
Publishing. IOP Conf. Series: Earth and Environmental Science, 5. https://doi.org/
Drinia, H., Antonarakou, A., Tsaparas, N., 2004. Diversity and abundance trends of
10.1088/1755-1307/5/1/012004.
benthic foraminifera from the southern part of the Iraklion Basin, Central Crete. Bull.
Harzhauser, M., Kroh, A., Mandic, O., Piller, W.E., Göhlich, U., Reuter, M., Berning, B.,
Geol. Soc. Greece 36, 772–781.
2007. Biogeographic responses to geodynamics: a key study all around the Oligo
Drinia, H., Antonarakou, A., Tsaparas, N., 2018. Diversity and abundance trends of
Miocene Tethyan Seaway. Zoologischer Anzeiger A Journal of Comparative Zoology
benthic foraminifera from the southern part of the Iraklion Basin, Central Crete. Bull.
246 (4), 241–256.
Geol. Soc. Greece 36 (2), 772–781.
Hayward, B.W., Le Coze, F., Gross, O., 2018. World Foraminifera Database. http://www.
Eichler, P.P.B., Felipe, M., Pimenta, F.M., Beatriz, B., Eichler, B.B., Helenice, Vital H.,
marinespecies.org/foraminifera.
2016. Living benthic foraminiferal species as indicators of cold-warm water masses
Herguera, J.C., Berger, W.H., 1991. Paleoproductivity from benthic foraminifera
interaction and upwelling areas. Cont. Shelf Res. 116, 116–121.
abundance: glacial to postglacial change in the West-equatorial Pacific. Geology 19,
Felix, R., 1973. Oligo-Miocene stratigraphy of Malta and Gozo. Meded Landbouwhogesch
1173–1176.
Wagening 73, 1–104.
Holbourn, A.E., Kuhnt, W., Schulz, M., Erlenkeuser, H., 2005. Impacts of orbital forcing
Fentimen, R., Rüggeberg, A., Lim, A., El Kateb, A., Foubert, A., Wheeler, A.J.,
and atmospheric CO2 on Miocene ice-sheet expansion. Nature 438, 483–487.
Spezzeferri, S., 2018. Benthic foraminifera in a deep-sea high-energy environment:

15
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

Holbourn, A., Kuhnt, W., Schulz, M., Flores, J.-A., Andersen, N., 2007. Orbitally-paced Melki, T., Kallel, N., Jorissen, F.J., Guichard, F., Dennielou, B., Berne, S., Labeyrie, L.,
climate evolution during the middle Miocene “Monterey” carbon-isotope excursion. Fontugne, M., 2009. Abrupt climate change, sea surfacesalinity and
Earth Planet. Sci. Lett. 261, 534–550. paleoproductivity in the western Mediterranean Sea(Gulf of Lion) during the last 28
Holbourn, A., Henderson, A.S., MacLeod, N., 2013a. Atlas of Benthic Foraminifera. kyr. Palaeogeogr. Palaeoclimatol. Palaeoecol. https://doi.org/10.1016/j.
Wiley-Blackwell and Natural History Museum, Oxford, UK, 642 pp. palaeo.2009.05.005.
Holbourn, A., Kuhnt, W., Clemens, S., Prell, W., Andersen, N., 2013b. Middle to late Milker, Y., Jorissen, F.J., Riller, U., Reicherter, K., Titschack, J., Weinkauf, M.F.G.,
Miocene stepwise climate cooling: evidence from a high-resolution deep water Theodor, M., Schmied, G., 2019. Paleo-ecologic and neotectonic evolution of a
isotope curve spanning 8 million years. Paleoceanography 28, 688–699. https://doi. marine depositional environment in SE Rhodes (Greece) during the early
org/10.1002/2013PA002538. Pleistocene. Quat. Sci. Rev. 213, 120–132.
Holbourn, A., Kuhnt, W., Kochhann, K.G.D., Andersen, N., Meier, K.J.S., 2015. Global Miller, K.G., Browning, J.V., Schmelz, W.J., Kopp, R.E., Mountain, G.S., Wright, J.D.,
perturbation of the carbon cycle at the onset of the Miocene Climatic Optimum. 2020. Cenozoic sea-level and cryospheric evolution from deep-sea geochemical and
Geology 43 (2), 123–126. continental margin records. Sci. Adv. 6 (eaaz1346), 1–15.
Hyde, H.P.T., 1955. Geology of the Maltese Islands. Lux Press, Malta, p. 135. Miller, K.G., Kent, D.V., 1987. Testing Cenozoic eustatic changes: the critical role of
Ied, I.M., Holcová, K., Abd-Elshafy, E., 2011. Biostratigraphy and paleoecology of the stratigraphic resolution. Cushman Found. Foraminiferal Res. Spec. Publ. 24, 51–56.
Burdigalian-Serravallian sediments in Wadi Sudr (Gulf of Suez, Egypt): comparison Miller, K.G., Wright, J.D., Fairbanks, R.G., 1991. Unlocking the ice house: Oligocene-
with the Central Paratethys evolution. Geol. Carpath. 62 (3), 233–249. https://doi. Miocene oxygen isotopes, eustasy and margin erosion. J. Geophys. Res. 96B,
org/10.2478/v10096-011-0019-6. 6829–6848.
Jorissen, F.J., 1999. Benthic foraminiferal microhabitats below the sediment-water Moodley, L., Hess, C., 1992. Tolerance of Infaunal Benthic Foraminifera for Low and
interface. In: Sen Gupta, B.K. (Ed.), Modern foraminifera. Kluwer Academic High Oxygen Concentrations. Biol. Bull. 183 (1), 94–98.
Publishers, pp. 161–179. Mourik, A.A., Abels, H.A., Hilgen, F.J., Di Stefano, A., Zachariasse, W.J., 2011. Improved
Jorissen, F.J., 2003. Benthic foraminiferal microhabitats below the sediment-water astronomical age constraints for the middle Miocene climate transition based on
interface. In: Sen Gupta, B.K. (Ed.), Modern foraminifera. Kluwer Academic high–resolution stable isotope records from the Central Mediterranean Maltese
Publishers, pp. 161–179. Islands. Paleoceanography 26, PA1210, 14 PP. https://doi.org/10.1029/2010PA001
Jorissen, F.J., Wittling, I., 1999. Ecological evidence from live–dead comparisons of 981.
benthic foraminiferal faunas off Cape Blanc (Northwest Africa). Palaeogeogr. Murray, J.W., 1991. Ecology and palaeoecology of benthic foraminifera. Longman,
Palaeoclim. 149, 151–170. Harlow.
Jorissen, F.J., De Stigter, H.C., Widmark, J.G.V., 1995. A conceptual model explaining Murray, J.W., 2003. An illustrated guide to the benthic foraminifera of the Hebridean
benthic foraminiferal microhabitats. Mar. Micropaleontol. 26, 3–15. shelf, west of Scotland, with notes on their mode of life. Palaeontol. Electron. 5 (1),
Jorissen, F.J., Wittling, I., Peypouquet, J.P., Rabouille, C., Relexans, J.C., 1998. Live 31.
benthic foraminiferal faunas off Cape Blanc, NW Africa; community structure and Murray, J.W., 2006. Ecology and applications of benthic foraminifera. Cambridge
microhabitats. Deep Sea Res. I 45, 2157–2188. University Press, p. 426.
Jorissen, F.J., Fontanier, C., Thomas, E., 2007. Paleoceanographical proxies based on Nigam, R., Mazumder, A., Henriques, P.J., Saraswat, R., 2007. Benthic foraminifera as
deep-sea benthic foraminiferal assemblage characteristics. In: Hillaire-Marcel, C., de proxy for oxygen-depleted conditions off the central West coast of India. J. Geol. Soc.
Vernal, A. (Eds.), Proxies in late Cenozoic Paleoceanography: Pt. 2: Biological tracers India 70, 1047–1054.
and biomarkers. Elsevier, pp. 263–326. Oblak, C., 2007. Most abundant middle Miocene rotaliinas (suborder Rotaliina,
Kaiho, K., 1994. Benthic foraminiferal dissolved-oxygen index and dissolved-oxygen Foraminifera) of Kozjansko (Eastern Slovenia). Geologija 50 (2). https://doi.org/
levels in the modern ocean. Geology 22, 719–722. 10.5474/geologija.2007.021.
Kaiho, K., 1999. Effect of organic carbon flux and dissolved oxygen on the benthic Oil Exploration Directorate, 1993. Geological map of the Maltese Island; Sheet 1, Malta.
foraminiferal oxygen index (BFOI). Mar. Micropaleontol. 37, 67–76. Keyworth: British Geological Survey. Publication of the Oil Exploration Directorate,
Karami, M.P., de Leeuw, A., Krijgsman, W., Meijer, P.Th., Wortel, M.J.R., 2011. The role Office of the Prime Minister, Malta (C. Simpson Cartographer).
of gateways in the evolution of temperature and salinity of semi-enclosed basins: an Okada, H., 1983. Modern nannofossil assemblages in sediments of coastal and marginal
oceanic box model for the Miocene Mediterranean Sea and Paratethys. Glob. Planet. seas along the Western Pacific Ocean. In: Meulenkamp, J.E. (Ed.), Reconstruction of
Chang. 79, 73–88. Marine Paleoenvironments, Utrecht Micropl. Bull, vol. 30, pp. 171–187.
Katz, M.E., Miller, K.G., 1993. (Table 1) Biostratigraphic datums of ODP Leg 133 sites. Pascual, A., Rodríguez-Lázaro, J., Martínez-García, B., Varela, Z., 2020.
PANGAEA, https://doi.org/10.1594/PANGAEA.785475, Supplement to: Katz M.E., Palaeoceanographic and palaeoclimatic changes during the last 37,000 years
Miller K.G., 1993. Neogene subsidence along the northeastern Australian Margin: detected in the SE Bay of Biscay based on benthic foraminifera. Quat. Int. 566-567,
benthic foraminiferal evidence. In: Mckenzie, J.A., Davies, P.J., Palmer-Julson, A., 323–336.
et al. (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results, College Patarroyo-Camargo, G.D., Martinez-Rodriguez, J.I., 2013. Deep-sea benthic foraminifera
Station, TX (Ocean Drilling Program), 133, pp. 75–92. In: https://doi.org/10.2973 of the Panama basin: ecology and their possible relation with deep-sea currents.
/odp.proc.sr.133.242.1993. Boletín de Investigaciones Marinas y Costeras 42 (1), 31–55.
Kouwenhoven, T.J., 2000. Survival under stress: benthic foraminiferal patterns and Pedley, H.M., 1976. A palaecological study of the Upper Coralline Limestone
Cenozoic biotic crises. Geol. Ultraiect. 186, 206. Terebratula–Aphelesia Bed (Miocene, Malta) based on bryozoan growth-forms and
Kouwenhoven, T.J., van der Zwaan, G.J., 2006. A reconstruction of late Miocene brachiopod distribution. Palaeogeogr. Palaeoclim. 20, 209–234.
Mediterranean circulation patterns using benthic foraminifera. Palaeogeogr. Pedley, H.M., 1978. A new lithostratigraphical and paleoenvironmental interpretation
Palaeoclim. 238, 373–385. for the coralline limestone formations (Miocene) of the Maltese Islands. Inst. Geol.
Levy, A., Mathieu, R., Poignant, A., Rosset-Moulinierd, M., Ambroise, D., 1995. Benthic Sci. Overseas Geol. Miner. Resour. 54, 273–291.
foraminifera from the Fernando de Noronha Archipelago (northern Brazil). Mar. Pedley, H.M., House, M.R., Waugh, B., 1976. The geology of Malta and Gozo. Proc. Geol.
Micropaleontol. 26, 89–97. As. 87 (3), 325–341.
Licari, L., Mackensen, A., 2005. Benthic foraminifera off West Africa (18Nto328S): do Pedley, H.M., House, M.R., Waugh, B., 1978. The geology of the Pelagian block: the
live assemblages from the topmost sediment reliably record environmental Maltese Islands. In: Nairn, A.E.M., Kanes, W.H., Stehli, F.G. (Eds.), The ocean basins
variability? Mar. Micropaleontol. 55, 205–233. and margins: the Western Mediterranean. Plenum Press, New York and London,
Lirer, F., Caruso, A., Cosentino, C., Turco, E., Sierro, F., Salvatorini, G., Foresi, L.M., pp. 417–433.
Iaccarino, S.M., 2019. Mediterranean Neogene planktonic foraminifer biozonation Pérez-Asensio, J.N., Aguirre, J., Schmiedl, G., Civi, S.J., 2012. Impact of restriction of the
and biochronology. Earth Sci. Rev. 196 https://doi.org/10.1016/j. Atlantic-Mediterranean gateway on the Mediterranean Outflow Water and eastern
earscirev.2019.05.013. Atlantic circulation during the Messinian. Paleoceanography 27, PA3222. https://
Loeblich, A.R., Tappan, H., 1964. Sarcodina chiefly “Thecamoebian” and Foraminiferida. doi.org/10.1029/2012PA002309.
In: Moore, R.C. (Ed.), Treatise on Invertebrate Paleontology, Partc C, Protista 2. Pérez-Asensio, J.N., Frigola, J., Pena, L.D., Sierro, F.J., Reguera, M.I., Rodríguez-
Lawrence: Geol. Soc. America Univ. Kansas press, p. 900. Tovar, F.J., Dorador, J., Asioli, A., Kuhlmann, J., Huhn, K., Cacho, I., 2020. Changes
Loeblich, A.R., Tappan, H., 1988. Foraminiferal genera and their Classification. Van in western Mediterranean thermohaline circulation in association with a deglacial
Nostrand Reinhold Comp, New York, p. 970. Organic Rich Layer formation in the Alboran Sea. Quat. Sci. Rev. 228, 106075.
Loubere, P., Fariduddin, M., 1999. Benthic foraminifera and the flux of organic carbon to Prista, G.A., Agostinho, R.J., Cachão, M.A., 2015. Observing the past to better
the seabed. In: Sen Gupta, B.K. (Ed.), Modern foraminifera. Kluwer Academic understand the future: a synthesis of the Neogene climate in Europe and its
Publishers, pp. 181–199. perspectives on present climate change. Open Geosci. 65–83.
Lutze, G.F., 1986. Uvigerina species of the Eastern North Atlantic. Utrecht Pucci, F., Geslin, E., Barras, C., Morigi, C., Sabbatini, A., Negri, A., Jorissen, F.J., 2009.
Micropaleontol. Bull. 35, 21–46. Utrecht. Survival of benthic foraminifera under hypoxic conditions: Results of an
Martins, V., Dubert, J., Jouanneau, J.-M., Weber, O., Ferreira, E., da Silva, E., Patinha, C., experimental study using the CellTracker Green method. Mar. Pollut. Bull. 59
Alveirinho Dias, J.M., Rocha, F., 2007. A multiproxy approach of the Holocene (8–12), 336–351.
evolution of shelf–slope circulation on the NW Iberian Continental Shelf. Mar. Geol. Rai, A.K., Singh, V.B., 2012. Response of eastern Indian Ocean (ODP Site 762B) benthic
239, 1–18. foraminiferal assemblages to the closure of the Indonesian seaway. Oceanologia 54
Mathieu, R., 1986. Sediments et Foraminifères actuels de la marge continentale (3), 449–472. https://doi.org/10.5697/oc.54-3.449.
Atlantique du Maroc. Thèse de Doctorat d’ Etat ès Sciencies Naturelles. Universitè Ramsay, A.T.S., Smart, C.W., Zachos, J.C., Cramp, A., Macleod, C.J., 1998. A model of
Pierre et Marie Curie, p. 420. early to middle Miocene deep ocean circulation for the Atlantic and Indian Oceans.
Mazzei, R., 1985. The Miocene sequence of the Maltese Islands: biostratigraphic and In: Lee, S.V., Jones, E.J.W. (Eds.), Geological evolution of ocean basins: results from
chronostratigraphic references based on nannofossils. Atti Soc. Tosc. Sci. Nat., Mem. the Ocean Drilling Program. Geol. Soc., London, Spec. Publ, p. 131.
Ser A 92, 165–197.

16
B. Russo et al. Palaeogeography, Palaeoclimatology, Palaeoecology 586 (2022) 110722

Rasmussen, T.L., Thomsen, E., 2017. Ecology of deep-sea benthic foraminifera in the Spötl, C., Vennemann, T.W., 2003. Continuous-flow isotope ratio mass spectrometric
North Atlantic during the last glaciation: food or temperature control. Palaeogeogr. analysis of carbonate minerals. Rapid Commun. Mass Spectrom. 17 (9), 1004–1006.
Palaeoclim. 472, 15–32. https://doi.org/10.1016/j.palaeo.2017.02.012. https://doi.org/10.1002/rcm.1010.
Rasmussen, T.L., Bäckström, D., Heinemeier, J., Klitgaard-Kristensen, D., Knutz, P.C., Sprovieri, M., Sgarrella, F., Russo, B., Bellanca, A., Neri, R., 2004. A Milankovitch climate
Kuijpers, A., Lassen, S., Thomsen, E., Troelstra, S.R., van Weering, T.C.E., 2002. The control on the middle Miocene Mediterranean Intermediate Water. In: D’Argenio, B.,
Faroe^Shetland Gateway: late Quaternary water mass exchange betweenthe Nordic Fisher, A.G., Premoli, Silva I., Weissert, H., Ferreri, V. (Eds.), Cyclostratigraphy:
seas and the northeastern Atlantic. Mar. Geol. 188, 165–192. Approaches and Case Histories, SEPM - Society for Sedimentary Geology, Special
Rathburn, A.E., Willingham, J., Ziebis, W., Burkett, A.M., Bruce, H., Corliss, B.H., 2018. Publication, vol. 81, pp. 45–55.
A New biological proxy for deep-sea paleo-oxygen: Pores of epifaunal benthic Stefanelli, S., 2004. Cyclic changes in oxygenation based on foraminiferal
foraminifera. Sci. Rep. 8, 9456. https://doi.org/10.1038/s41598-018-27793-4. michrohabitats: early-middle Pleistocene, Lucania Basin (southern Italy).
Rehault, J.P., Boillot, G., Mauffret, A., 1985. The western Mediterranean Basin. In: J. Micropaleontol. 23, 81–95, 0262-821X/04.
Stanley, D.J., Wezel, F.C. (Eds.), Geological evolution of the Mediterranean Basin. Steininger, F.F., Rabeder, G., Rogl, F., 1985. Land mammal distribution in the
Springer - Verlag, New York, pp. 101–130. Mediterranean Neogene: a consequence of geokinematic and climatic events. In:
Rizzo, C., 1932. Report on the Geology of the Maltese Islands. Government Printing Stanley, D.J., Wezel, F.C. (Eds.), Geological evolution of the Mediterranean Basin.
Office, Malta, p. 37. Springer - Verlag, New York, pp. 559–571.
Rögl, F., Spezzaferri, S., 2003. Foraminiferal paleoecology and biostratigraphy of the Steinthorsdottir, M., Coxall, H.K., de Boer, A.M., Huber, M., Barbolini, N., Bradshaw, C.
Mühlbach section (Gaindorf Formation, lower Badenian), Lower Austria. Annalen D., Burls, N.J., Feakins, S.J., Gasson, E., Henderiks, J., Holbourn, A.E., Kiel, S.,
des naturhistorischen Museums in Wien, Serie a 104, 23–75. Kohn, M.J., Knorr, G., Kürschner, W.M., Lear, C.H., Liebrand, D., Lunt, D.J., Mörs, T.,
Rögl, F., Steininger, F.F., 1983. Vom zerfall der Tethys zu Mediterran und Paratethys, die Pearson, P.N., Pound, M.J., Stoll, H., Strömberg, C.A.E., 2021. The Miocene: the
Neogene Palaeographie und Palinspastik des zirkum-Mediterranen Raumes. Ann. future of the past. Paleoceanogr. Paleoclimatol. 36 https://doi.org/10.1029/
Naturhist. Mus. Wien 85 (A), 135–163. 2020PA004037.
Russo, B., Sgarrella, F., Gaboardi, S., 2002. Benthic foraminifera as indicators of Székely, S.F., Bindiu-Haitonic, R., Filipescu, S., Bercea, R., 2017. Biostratigraphy and
palaecological bottom conditions in the Serravallian Tremiti sections (eastern paleoenvironmental reconstruction of the marine lower Miocene Chechiș Formation
Mediterranean, Italy). Riv. It. Paleontol. S. 108 (2), 275–287. in the Transylvanian Basin based on foraminiferal assemblages. Carnets Geol Madrid
Russo, B., Curcio, E., Iaccarino, S., 2007. Paleoecology and paleoceanography of a 17 (2), 11–37.
Langhian succession (Tremiti Islands, southern Adriatic Sea, Italy) based on benthic Thomas, E., 2007. Cenozoic mass extinctions in the deep sea; what disturbs the largest
foraminifera. Boll. Soc. Paleontol. It. 46 (2–3), 107–124. habitat on Earth? In: Monechi, S., Coccioni, R., Rampino, M. (Eds.), Large Ecosystem
Sander Ernst, S., van der Zwaan, B., 2004. Effects of experimentally induced raised levels Perturbations: Causes and Consequences, Geol. S. Am. S, vol. 424, pp. 1–24.
of organic flux and oxygen depletion on a continental slope benthic foraminiferal Turco, E., Hüsing, S., Hilgen, F., Cascella, A., Gennari, R., Iaccarino, S.M., Sagnotti, L.,
community. Deep-Sea ResPT I 51, 1709–1739. 2017. Astronomical tuning of the La Vedova section between 16.3 and 15.0 Ma.
Sarkar, S., Gupta, A.K., 2009. Late Quaternary benthic foraminifera from Ocean Drilling Implications for the origin of megabeds and the Langhian GSSP. Newsl. Stratigr. 50,
Program Hole 716A, Maldives Ridge, southeastern Arabian Sea. Micropaleontology 1–29.
55 (1), 1–11. Van der Zwaan, G.J., Duijnstee, I.A.P., Den Dulk, M., Ernst, S.R., Jannink, N.T.,
Schmiedl, G., Mackensen, A., Muller, P., 1997. Recent benthic foraminifera from the Kouwenhoven, T.J., 1999. Benthic foraminifers: proxies or problems? A review of
eastern South Atlantic Ocean: dependence of food supply and water masses. Mar. palaecological concepts. Earth Sci. Rev. 46, 213–236.
Micropaleontol. 32, 249–288. Van Hinsbergen, D.J.J., Kouwenhoven, T., van der Zwaan, G.J., 2005. Paleobathymetry
Schmiedl, G.F., de Bove, F., Buscailc, R., Charriére, B., Hemleben, C., Medernach, L., in the backstripping procedure: Correction for oxygenation effects on depth
Piconc, P., 2000. Trophic control of benthic foraminiferal abundance and estimates. Palaeogeogr. Palaeoclimatol. Palaeoecol. 221, 245–265.
microhabitat in the bathyal Gulf of Lions, western Mediterranean Sea. Mar. Van Morkhoven, F.P.C.M., Berggren, W.A., Edwards, A.S., 1986. Cenozoic cosmopolitan
Micropaleontol. 40, 167–188. deep-water benthic foraminifera. In: Oertli, H.J. (Ed.), Bulletin des Centres de
Schmiedl, G., Mitschele, A., Beck, S., Emeis, K.-C., Hemleben, C., Schulz, H., Sperling, M., Recherches Exploration-Production, Elf-Aquitaine Memory, 11. Elf-Aquitaine, Pau,
Weldeab, S., 2003. Benthic foraminiferal record of ecosystem variability in the p. 421.
eastern Mediterranean Sea during times of sapropel S5 and S6 deposition. Verducci, M., Foresi, L.M., Scott, G.H., Sprovieri, M., Lirer, F., Pelosi, N., 2009. The
Palaeogeogr. Palaeoclimatol. Palaeoecol. 190, 139–164. Middle Miocene climatic transition in the Southern Ocean: evidence of paleoclimatic
Seiglie, G.A., 1968. Foraminiferal assemblages as indicators of organic carbon content or and hydrographic changes at Kerguelen plateau from planktonic foraminifers and
of water pollution. Am. Assoc. Pet. Geol. 52, 2231–2241. stable isotopes. Palaeogeogr. Palaeoclimatol. Palaeoecol. 280 (3–4), 371–386.
Sen Gupta, B.K., Machain-Castillo, M.L., 1993. Benthic foraminifera in oxygen-poor Verhallen, P.J.J.M., 1991. Late Pliocene to early Pleistocene Mediterranean mud-
habitats. Mar. Micropaleontol. 20, 183–201. dwelling foraminifera; influence of a changing environment on community structure
Sgarrella, F., Di Donato, V., Sprovieri, R., 2012. Benthic foraminiferal assemblage and evolution. Utrecht Micropaleontol. Bull. 40, 1–219.
turnover during intensification of the Northern Hemisphere glaciation in the Violanti, D., Bonci, M.C., Trenkwalder, S., Lozar, F., Beccaro, P., Dela Pierre, F.,
Piacenzian Punta Piccola section (Southern Italy). Palaeogeogr. Palaeoclimatol. Bernardi, E., Boano, P., 2011. Micropalaeontological evidences of high productivity
Palaeoecol. 333-334, 59–74. episodes in the Zanclean of Piedmont (Early Pliocene, northwestern Italy). Boll. Soc.
Shackleton, N.J., Kennett, J.P., 1975. Paleotemperature history of the Cenozoic and the Pal. It. 50 (2), 111–133. https://doi.org/10.4435/BSPI.2011.12.
initiation of Antarctic glacia-tion: Oxygen and carbon isotopic analyses in DSDP Sites Woodruff, F., Savin, S.M., 1989. Miocene deepwater oceanography. Paleoceanography 4,
277, 279, and 281. Initial Rep. Deep Sea Drill. Proj. 29, 143–156. 87–140.
Singh, A.D., Rai, A.K., Tiwari, M., Naidu, P.D., Verma, K., Chaturvedi, M., Niyogi, A., Woodruff, F., Savin, S., 1991. Mid-Miocene isotope stratigraphy in thedeep sea: High
Pandey, D., 2015. Fluctuations of Mediterranean Outflow Water circulation in the resolution correlations, paleoclimatic cycles, and sediment preservation.
Gulf of Cadiz during MIS 5 to 7: evidence from benthic foraminiferal assemblage and Paleoceanography 6, 755–806.
stable isotope records. Glob. Planet. Chang. 133, 125–140. Xue, L., Ding, X., Pei, R., Wan, X., 2019. Miocene paleoenvironmental evolution based on
Smart, C.W., Ramsay, A.T.S., 1995. Benthic foraminiferal evidence for the existence of an benthic foraminiferal assemblages in the Lufeng Sag, northern South China Sea. Acta
early Miocene oxygen-depleted oceanic water mass? J. Geol. Soc. Lond. 152 (5), Oceanol. Sin. 38 (3), 124–137.
735–738. Yasuda, N., 1997. Late Miocene-Holocene paleoceanography of the western equatorial
Spezzaferri, S., Ćorić, S., Hohenegger, J., Rögl, F., 2002. Basin-scale paleobiogeography Atlantic: evidence from deep-sea benthic foraminifers. Proceed. Ocean Drilling
and paleoecology: an example from Karpatian.(latest Burdigalian) benthic and Progr. Sci. Res. 154, 395–432.
planktonic foraminifera and calcareous nannofossils from the Central Paratethys. Zachos, J., Pagani, M., Sloan, L., Thomas, E., Billups, K., 2001. Trends, rhythms and
Geobios. Mémoire Spécial 24, 241–256. aberrations in global climate 65Ma to present. Science 292, 686–693.
Spezzaferri, S., Rüggeberg, A., Stalder, C., Margreth, S., 2013. Benthic foraminifer Zachos, J.C., Dickens, G.R., Zeebe, R.E., 2008. An early Cenozoic perspective on
assemblages from norwegian cold-water coral reefs. J. Foraminifer. Res. 43 (1), greenhouse warming and carbon-cycle dynamics. Nature 451, 279–283. https://doi.
21–39. org/10.1038/nature06588.
Spezzaferri, S., Rüggeberg, A., Stalder, C., Margreth, S., 2014. Benthic Foramniferal
Assemblages from Cold-Water Coral Ecosystems. Cushman Foundation Special
Publication, No. 44, pp. 20–48.

17

You might also like