Hostname: page-component-7bb8b95d7b-cx56b Total loading time: 0 Render date: 2024-10-04T03:58:13.912Z Has data issue: false hasContentIssue false

Characterization of phreatic overgrowths on speleothems precipitated in the northern Adriatic during a sea-level stillstand at ca. 2.8 ka

Published online by Cambridge University Press:  15 January 2024

Nina Lončar*
Affiliation:
University of Zadar, Department of Geography, Trg kneza Višeslava 9, 23000 Zadar, Croatia
Sanja Faivre
Affiliation:
University of Zagreb, Faculty of Science, Department of Geography , Marulićev trg 19/II, Zagreb, Croatia
Blaž Miklavič
Affiliation:
University of Guam, Water and Environmental Research Institute of the Western Pacific (WERI), UOG Station, 303 University Dr, Mangilao, 96913, Guam
Bogdan P. Onac
Affiliation:
University of South Florida, School of Geosciences, 4202 East Fowler Avenue, NES 107, Tampa, USA
Victor J. Polyak
Affiliation:
University of New Mexico, Earth & Planetary Sciences, 221 Yale Blvd NE, NM 87131, Albuquerque, USA
Yemane Asmerom
Affiliation:
University of New Mexico, Earth & Planetary Sciences, 221 Yale Blvd NE, NM 87131, Albuquerque, USA
*
Corresponding author: Nina Lončar; Email: <[email protected]>
Rights & Permissions [Opens in a new window]

Abstract

We examined a Late Holocene sea-level stillstand using phreatic overgrowths on speleothems (POS) recovered from Medvjeđa Špilja [Bear Cave] (northern Adriatic Sea) from −1.28 ± 0.15 m below present mean sea level. Different mineralogical analyses were performed to characterize the POS and better understand the mechanisms of their formation. Results reveal that the fibrous overgrowth is formed of calcite and that both the supporting soda straw and the overgrowth have very similar trace element compositions. This suggests that the drip-water and groundwater pool from which the POS formed have similar chemical compositions. Four subsamples were dated by means of uranium-series. We found that ca. 2800 years ago, the relative sea level was stable for about 300 years at a depth of approximately −1.28 ± 0.15 m below the current mean sea level. This finding roughly corresponds with the end of a relatively stable sea-level period, between 3250 and 2800 cal yr BP, previously noted in the southern Adriatic. Our research confirms the presence of POS in the Adriatic region and establishes the Medvjeđa Špilja pool as a conducive environment for calcite POS formation, which encourages further investigations at this study site.

Type
Thematic Set: Speleothem Paleoclimate
Copyright
Copyright © The Author(s), 2024. Published by Cambridge University Press on behalf of Quaternary Research Center

INTRODUCTION

In paleo sea-level research, various indicators, including sedimentological, geomorphological, archaeological, biological, and historical sources are used and often combined (e.g., Faivre and Fouache, Reference Faivre and Fouache2003; Faivre et al., Reference Faivre, Bakran-Petricioli, Horvatinčić and Sironić2013; Shennan et al., Reference Shennan, Long, Benjamin, Horton, Shennan, Long, Benjamin and Horton2015). In coastal caves, hiatuses in speleothem growth signal a switch between vadose and phreatic conditions. Gascoyne et al. (Reference Gascoyne, Benjamin, Schwarcz and Ford1979) and Li et al. (Reference Li, Lundberg, Dickin, Ford, Schwarcz, McNutt and Williams1989) are among the first researchers who documented mineralogical changes on the surfaces of such hiatuses. Furthermore, investigations on previously submerged speleothems that contain marine biogenic overgrowths and marine boring organisms (e.g., Alessio et al., Reference Alessio, Allegri, Antonioli, Belluomini, Ferranti, Importa, Manfra and Proposito1992, Reference Alessio, Allegri, Antonioli, Belluomini, Improta, Manfra and Preite Martinez1994; Antonioli and Oliverio, Reference Antonioli and Oliverio1996) provide an additional tool to assess past sea-level positions in littoral caves (Onac et al., Reference Onac, Ginés, Ginés, Fornós, Dorale, Ginés, Ginés, Gómez-Pujol, Onac and Fornós2012; van Hengstum et al., Reference van Hengstum, Richards, Onac, Dorale, Shennan, Long and Horton2015). The study of submerged speleothems in sea-level reconstructions has contributed significantly to the understanding of regional and global sea-level changes, especially for the western (Antonioli et al., Reference Antonioli, Cremona, Immordino, Puglisi, Romagnoli, Silenzi, Valpreda and Verrubbi2002, Reference Antonioli, Bard, Potter, Silenzi and Improta2004a, Reference Antonioli, Furlani, Montagna and Stocchi2021; Bard et al., Reference Bard, Antonioli and Silenzi2002; Stocchi et al., Reference Stocchi, Antonioli, Montagna, Pepe, Lo Presti, Caruso and Corradino2017) and the eastern Mediterranean basin (Surić et al., Reference Surić, Juračić, Horvatinčić and Krajcar Bronić2005, Reference Surić, Richards, Hoffmann, Tibljaš and Juračić2009; Surić and Juračić, Reference Surić and Juračić2010). Based on the age of marine overgrowth on speleothems, segments of the relative sea-level curve for the last 220 ka, which have been constructed for the eastern Adriatic coast (Surić and Juračić, Reference Surić and Juračić2010), are in general agreement with the global sea-level curve. The Early Holocene sea-level rise reached −41.5 m at ca. 9.2 ka and −10 m at ca. 7.8 ka and rose to −1.5 m by ca. 3.4 ka (Surić and Juračić, Reference Surić and Juračić2010).

Sea-level studies based on phreatic overgrowths on speleothems (POS) have been conducted since the early 1970s (Ginés and Ginés, Reference Ginés and Ginés1974). Unlike the submerged speleothems and biogenic encrustations, POS allow precise sea-level reconstructions (Vesica et al., Reference Vesica, Tuccimei, Turi, Fornós, Ginés and Ginés2000; Tuccimei et al., Reference Tuccimei, Soligo, Ginés, Ginés, Fornós, Kramers and Villa2010; Polyak et al., Reference Polyak, Onac, Fornós, Hay, Asmerom, Dorale, Ginés, Tuccimei and Ginés2018; Onac et al., Reference Onac, Mitrovica, Ginés, Asmerom, Polyak, Tuccimei and Ashe2022). POS are secondary depositional structures (carbonate encrustations) that precipitate in coastal caves at the water table around pre-existing vadose speleothems in favorable geochemical conditions (Ginés et al., Reference Ginés, Ginés, Pomar and Beck1981; Fornós et al., Reference Fornós, Gelabert, Ginés, Ginés, Tuccimei and Vesica2002). POS grow at sea level and within the tidal range for as long as sea level remains at the same elevation (Dumitru et al., Reference Dumitru, Polyak, Asmerom and Onac2021) (Fig. 1). These overgrowths are composed of aragonite and/or calcite, with the latter being more common. Each POS has an exact geographic location, its elevation can be measured with high precision, its morphology provides an indicative meaning (mean sea level), and it is datable by uranium-series method. These characteristics make POS ideal sea-level index points (van de Plassche, Reference van de Plassche1986), and thus excellent markers of sea-level change with local and global significance.

Figure 1. Conceptual model of phreatic overgrowths on speleothem (POS) formation (after Dumitru et al., Reference Dumitru, Polyak, Asmerom and Onac2021); msl = mean sea level.

In Mediterranean littoral caves (within ~300 m from the coastline), the hydraulic gradient between location and the sea is insignificant, so the cave water table is coincident with sea level, and was in the past (Dorale et al., Reference Dorale, Onac, Fornós, Ginés, Ginés, Tuccimei and Peate2010). Uranium-series (U-series) dating has shown that POS normally behave as closed systems, thus providing reliable ages (Tuccimei et al., Reference Tuccimei, Ginés, Delitala, Ginés, Gràcia, Fornós and Taddeucci2006, Reference Tuccimei, Van Strydonck, Ginés, Ginés, Soligo, Villa and Fornós2011; Dorale et al., Reference Dorale, Onac, Fornós, Ginés, Ginés, Tuccimei and Peate2010). This allows precise constraint to be placed on the timing of sea-level change, assuming that sea level remained at the same elevation for ca. 300 years or more (Polyak et al., Reference Polyak, Onac, Fornós, Hay, Asmerom, Dorale, Ginés, Tuccimei and Ginés2018; Dumitru et al., Reference Dumitru, Polyak, Asmerom and Onac2021).

According to Ginés et al. (Reference Ginés, Ginés, Fornós, Tuccimei, Onac, Gràcia, Ginés, Ginés, Gómez-Pujol, Onac and Fornós2012), the first description of carbonate encrustations (later defined as POS) refers to speleothems from Coves del Drac (Mallorca) by Rodés (Reference Rodés1925) and de Joly (Reference de Joly1929), who assumed that their formation was related to drowning events associated with past water-table elevations. The first thorough studies of POS as sea-level indicators in littoral caves of Mallorca began in 1972 when Ginés and Ginés (Reference Ginés and Ginés1974) proposed to relate subaqueous crystallization of speleothems from Cova de sa Bassa Blanca with past Pleistocene sea stands. Besides Mallorca, POS have been used for sea-level reconstructions in Sardinia (Tuccimei et al., Reference Tuccimei, Onac, Dorale, Ginés, Fornós, Ginés, Spada, Ruggieri and Mucedda2012), Japan (Miklavič et al., Reference Miklavič, Yokoyama, Urata, Miyairi and Kan2018), and Cuba (De Waele et al., Reference De Waele, D'Angeli, Tisato, Tuccimei, Soligo, Ginés and Ginés2017, Reference De Waele, D'Angeli, Bontognali, Tuccimei, Scholz, Jochum and Columbu2018). POS also have been identified and described in Bermuda (Harmon et al., Reference Harmon, Schwarcz and Ford1978) and Mexico (Jenson et al., Reference Jenson, Schwartz, Li and Gao2018). So far, POS-based research has contributed greatly to the study of sea level, in particular to precisely characterize the Late Pleistocene highstands, and to improve the glacial isostatic adjustments for the western Mediterranean (Tuccimei et al., Reference Tuccimei, Onac, Dorale, Ginés, Fornós, Ginés, Spada, Ruggieri and Mucedda2012; Polyak et al., Reference Polyak, Onac, Fornós, Hay, Asmerom, Dorale, Ginés, Tuccimei and Ginés2018; Onac et al., Reference Onac, Mitrovica, Ginés, Asmerom, Polyak, Tuccimei and Ashe2022). However, some POS that formed during the Late Holocene (Tuccimei et al., Reference Tuccimei, Soligo, Ginés, Ginés, Fornós, Kramers and Villa2010, Reference Tuccimei, Van Strydonck, Ginés, Ginés, Soligo, Villa and Fornós2011; Miklavič et al., Reference Miklavič, Yokoyama, Urata, Miyairi and Kan2018) and, during the last 2800 years in particular, show sea-level stability throughout that period (Onac et al., Reference Onac, Mitrovica, Ginés, Asmerom, Polyak, Tuccimei and Ashe2022).

Considering the abundance of karst forms in Croatia, including anchialine caves (Surić et al., Reference Surić, Lončarić and Lončar2010), one of the goals of the SEALeveL project (HRZZ IP-2019-04-9445) was to find POS in the Adriatic that would enable more robust relative sea-level change studies and to combine these results with results from other markers. The research presented herein is based on mineralogical and U-series analyses of the first POS discovered in the Adriatic Sea. Our aim is to characterize the POS and the environment of its formation, and to define the period of relative the sea-level stability related to its growth.

STUDY SITE

POS were found in Medvjeđa Špilja [Bear Cave] on Lošinj Island (Kvarner region, northern Adriatic; Fig. 2), which is located in the complex contact zone between Adriatic foreland, Istrian Peninsula, and the external Dinarides (Korbar, Reference Korbar2009; Schmid et al., Reference Schmid, Fügenschuh, Kounov, Matenco, Nievergelt, Oberhänsli and Pleuger2020; van Hinsbergen et al., Reference van Hinsbergen, Torsvik, Schmid, Matenco, Maffione, Vissers, Gürer and Spakman2020). According to Špelić et al. (Reference Špelić, Del Ben and Petrinjak2021), this area is dominated by an alternation of structural lows and highs, mainly oriented N–S, NNW–SSE, and NW–SE (Fig. 2). A comprehensive overview of the eastern Adriatic tectonic setting is given in Korbar (Reference Korbar2009). The area is mainly composed of Carboniferous to Eocene carbonate rocks that were deposited in shallow-marine environments (Vlahović et al., Reference Vlahović, Tišljar, Velić and Matičec2005). Prevalent strata on Lošinj Island are carbonate deposits of Cretaceous age, as well as Eocene foraminiferal limestone and Quaternary loess deposits (Korbar, Reference Korbar2009).

Figure 2. Study site and geologic setting (based on the Geological Map of the Republic of Croatia, scale 1:300 000, Croatian Geological Survey); MNE = Montenegro.

Formation of caves in karst is predominantly controlled by structural characteristics of the area, as is the case with Medvjeđa Špilja (Fig. 2b). Tectonics and favorable climatic conditions, along with Pliocene and Pleistocene sea-level fluctuations, caused carbonate areas to emerge, leading to the karstification of today's eastern Adriatic coast (Surić et al., Reference Surić, Lončarić and Lončar2010, Reference Surić, Korbar and Juračić2014). During the Late Pleistocene–Holocene marine transgression, karst features such as caves became submerged. The speleothems within these caves now provide a potential record for reconstructing Quaternary relative sea-level changes.

Medvjeđa Špilja is an anchialine cave developed in Cretaceous limestone and situated in the central part of the Lošinj Island, ~55 m from the sea and 17.5 m asl (Fig. 1). The entrance to the cave is through a narrow opening formed along a vertical fissure that extends perpendicular to the coast (Malez and Božičević, Reference Malez, Božičević and Stelcl1965). The entrance leads to a bell-shaped chamber with a lake at the bottom. The rest of the cave is a mostly submerged channel with a total length of 245 m stretching along a NNE–SSW trending fissure (Jalžić, Reference Jalžić2007) parallel to the coast (Fig. 3) The cave is linked to the sea through karstified fractures. The connection between fresh groundwater and the sea is rather direct, as often documented along the eastern Adriatic coast (Bonacci and Roje-Bonacci, Reference Bonacci and Roje-Bonacci2003). Short time in-situ measurements (see Methods) revealed a tidal range of 44 cm, whereas the long-term average in the area is 48 cm (Faivre et al., Reference Faivre, Pahernik and Maradin2011b). The existence of cave bear (Ursus spelaeus) remains (Malez et al., Reference Malez, Sliepčević and Srdoč1979) and collapse material in the seaward part of the cave, as observed in recent diving explorations, suggest an open horizontal connection to the coast in the past. The cave is rich in speleothems, which are mostly submerged. Salinity increases with depth. The halocline was observed at a depth of approximately −2 m during dive prospecting.

Figure 3. Medvjeđa Špilja (Bear Cave) longitudinal profile. (a, b) Cross-sectional and plan maps, (c) position in relation to the sea.

MATERIAL AND METHODS

Sampling and depth measurement

We conducted detailed cave diving explorations in March and October 2021, resulting in the recovery of suitable speleothems from all submerged parts of the cave. Sample MLp1 is a stalactite found at Little Lake passage rooftop (Fig. 3a) and collected in growth position at the uppermost part, at a depth of −1.28 ± 0.15 m. The speleothem is ~8 cm long, tapering from 0.5 cm at the top to ~2.0 cm at the base. This widening towards the base gives the speleothem a rounded shape, which is consistent with the morphology of POS. Its surface is light yellowish (Fig. 4). After collection, the sample was cut in half by diamond disc attached to a rotary drill mounted to a special constructed slider.

Figure 4. (a, b) External morphology and longitudinal section of phreatic overgrowth on speleothems (POS) sample MLp1, showing locations of samples for U-series dating (black marks); note the ~1-mm thick overgrowth layer immediately around the support (a pre-existing stalactite). (c) XRF spot analyses (red) on the halved and outer surface of the POS.

Sampling depth was measured on several occasions with the pressure depth-meter built in the Suunto EON steel diving computer, which has an accuracy of ± 1% and resolution of ± 0.1 m. All measurements revealed the same depth. To establish the depth of measurements, the uppermost part of the soda straw was chosen as reference point. The elevation of POS paleo-levels presented here is the sum of the depth at which the sample was taken and length of the sample, referenced to the current mean sea level (MSL).

To confirm the connection between groundwater and open sea, as well as the tidally induced oscillation within the cave pool, both sea-level and groundwater-level fluctuation were recorded using a HOBO U20-001-02 TI water level logger. The device measures pressure and converts it to water elevation using the HOBOware-Pro software package with a typical accuracy of 0.3 cm (Onset, 2022). The logger was deployed during the spring tide period between October 8 and October 11. One logger was placed inside the cave in Little Lake (Fig. 3), whereas the second one was placed in the sea at same depth. A high-resolution (10-min) recording was set in order to eliminate the possibility of a false signal caused by waves. The measurements recorded a spring tide range of 44 cm, and the high tide occurred on 9 October 2022 at 20:55 (CMHS, 2022). To account for atmospheric pressure, we used data from the Croatian Meteorological and Hydrological Service Weather Station of Mali Lošinj.

X-ray diffraction (XRD) analysis

The mineralogy of the phreatic overgrowths was determined by XRD. The analysis was performed at the Department of Geology, Faculty of Natural Sciences and Engineering, University of Ljubljana (FNSE-UL), Slovenia, using a Philips X-ray diffractometer generator PW 3830. About 1–3 g of sample was drilled from the outer part of the POS (Fig. 4) with a dental drill and subsequently crushed in an agate mortar.

Scanning electron microscopy–energy dispersive spectroscopy (SEM–EDS) analyses

The analyses focused on characterizing the structural and chemical differences between the supporting soda straw and overgrowth at the macroscopically visible boundary between the two units. Analyses were performed at FNSE-UL by ThermoFisher Scientific Quattro S with Schottky effect field-emission gun SEM (FEG–SEM) with an Oxford Instruments UltimMax 65 energy-dispersive spectrometer (EDS) on the polished surface of the POS longitudinal section. Structural etching (Herwegh, Reference Herwegh2000) was used to reveal the fabric structure and helped distinguish this boundary on SEM images. Elemental mapping and spot analyses were used for chemical characterization of the aforementioned parts of the POS. Spot analyses targeted individual crystals on both sides of the boundary.

X-ray fluorescence (XRF) analysis

XRF analysis aimed to geochemically characterize the POS and to determine if there is any difference in trace element composition between the support and the overgrowth. Special attention was paid to Mg since it was reported in previous studies to be higher in overgrowths (Vesica et al., Reference Vesica, Tuccimei, Turi, Fornós, Ginés and Ginés2000; Ginés et al., Reference Ginés, Fornós, Ginés, Sanjaume and Mateu2005; Csoma et al., Reference Csoma, Goldstein and Pomar2006). MLp1 was analyzed with a Thermo Scientific Niton XL5+ XRF instrument having a 3-mm analyzing spot size at FNSE-UL. Nearly pure and partly dolomitized limestone standards (NIST-1d and NIST-88b, respectively) were used for calibration to obtain good accuracy of trace elements in a CaCO3 matrix. We conducted three spot analyses on the support (#1–3), and five spot analyses on the overgrowth (#4–8; three on the cut surface [#4–6] and two on the outer surface [#7, #8] where the natural surface was abraded off by a dental drill—the resulting powder was used for XRD analysis) (Fig. 4).

U-series dating

To obtain the deposition time of our POS, four subsamples were taken across the thickest part of the phreatic overgrowth. To document the age when the cave became submerged, the first subsample was drilled 1 mm away from the pre-existing soda straw. The second and third subsamples are located between the vadose soda straw and POS's external surface, while the last subsample comes from the outermost part of the POS (Fig. 4). The pre-existing soda straw was also dated. All subsamples were dated by means of U-series disequilibrium method on a Thermo Neptune Multi-Collector Inductively Coupled Plasma Mass Spectrometer (MC-ICPMS) at the Department of Earth and Planetary Sciences, University of New Mexico, USA. Details on this method are available in Asmerom et al. (Reference Asmerom, Polyak, Schwieters and Bouman2006).

Since all four POS subsample dates are within a few hundred years of each other (using the often-assumed atomic ratio value of initial 230Th/232Th to be 4.4 ppm based on the bulk Earth 232Th/238U value of 3.8 [Cheng et al., Reference Cheng, Edwards, Shen, Polyak, Asmerom, Woodhead and Hellstrome2013]), we calculated 232Th/238U–234U/238U–230Th/238U isochron age that represents a more robust overall age for the POS. The isochron was constructed using IsoplotR (Vermeesch, Reference Vermeesch2018), which yielded a measured initial 230Th/232Th atomic ratio of 9.7 ± 0.86 ppm, which is double the value traditionally used in calculating U-series ages (4.4 ± 2.2 ppm). With the measured initial 230Th/232Th atomic ratio value, we re-calculated all four POS ages (at 1, 4, 6, and 8 mm). The zero datum for all ages is AD 1950 and all ages are reported with absolute 2σ uncertainty (Table 1). Then we weight-averaged those ages and uncertainties to produce an overall robust time of deposition for the POS.

Table 1. Summary of the U-series measurement results for the phreatic overgrowth of speleothems sample MLp1.

Subsample powder sizes range from 60 to 120 mg; 1–8 mm labels reflect the distance from the stalactite outer surface. Initial 230Th/232Th atomic ratio used to correct ages is 0.0000097 (activity ratio = 1.8) ± 10% based on a 4-point isochron. The 232Th/238U–234U/238U–230Th/238U isochron age = 2.80 ± 0.09 yr was calculated using IsoplotR (Vermeesch, Reference Vermeesch2018). All errors are absolute 2σ. AR = activity ratio; δ234Um = measured value and δ234Ui = initial value. Zero datum for all ages is AD 1950. Weight average is calculated from the four POS samples.

RESULTS

Diving expeditions resulted in discovery of the first POS located in the Adriatic Sea. The sample MLp-1 is an 8-cm long calcite soda straw with a phreatic overgrowth (Fig. 4). The overgrowth has an uneven deposition pattern over a regularly shaped soda straw, widening towards the bottom and terminating in a rounded base (Fig. 4), a feature commonly observed in Mallorcan caves (Ginés et al., Reference Ginés, Ginés, Fornós, Tuccimei, Onac, Gràcia, Ginés, Ginés, Gómez-Pujol, Onac and Fornós2012). An ~1 mm thick darker overgrowth layer immediately around the pre-existing support is easily visible (Fig. 4).

XRD, XRF, and SEM–EDS analyses

Calcite was the only mineral detected phase in the POS. The etched surface revealed the shapes of the crystals. The support crystals tend to be smaller and etch differently than the overgrowth crystals, which gives them a fuzzier appearance (Fig. 6a). The overgrowth crystals often show epitaxial growth on top of the soda straw support crystals (Fig. 6b), which differs from previously observed boundaries between the support and the overgrowth (Ginés et al., Reference Ginés, Ginés, Fornós, Tuccimei, Onac, Gràcia, Ginés, Ginés, Gómez-Pujol, Onac and Fornós2012; Miklavič et al., Reference Miklavič, Yokoyama, Urata, Miyairi and Kan2018). Zoning can be observed in both support and overgrowth crystals, although it is more common in overgrowth crystals. The ~1-mm thick, dark overgrowth layer that is easily visible with the naked eye (Fig. 4) is much more difficult to identify on SEM images. The crystals are larger and more elongated with the long crystallographic c-axis being perpendicular to the support surface away from the boundary. Macroscopically, this gives a fibrous appearance to the crystals. The results of XRF analysis indicate that the concentration of Mg in calcite is equally low in the support and the overgrowth (Fig. 5), ranging between 0.44 and 0.62 wt%. Other detected elements (Si, Al, Fe, Sr, and Ba) also were present in all sampled areas of the POS and, just as Mg, they did not show any clear trend in their distribution across the POS (Fig. 6c).

Figure 5. Trace element content in the analyzed spots indicated in Figure 4 (red circles).

Figure 6. (a) The three parts of the POS as seen macroscopically (left) and under SEM (rigth; the area is indicated by white rectangle on the macroscopic picture). (b) The boundary between the support and the overgrowth with visible epitaxial crystal overgrowth in the inset picture. (c). SEM image showing the boundary (red dashed line) between the support and the dark overgrowth layer, and the EDS spot analysis at the support–overgrowth boundary (1 mm dark layer), showing Mg content. Observe the crystal zoning near spot #3. (d) EDS elemental map of the area shown in (a) showing a uniform distribution of Mg in the calcite across all three layers, and localized concentration of Si and Al.

EDS elemental mapping and spot analyses showed the presence of Mg, Si, Al, Na, Cl, and S. Mg content is uniformly distributed across the POS. The occurrence of Si in clusters or associated with Al, as shown on elemental maps, indicates that these elements are related to quartz and clay particles incorporated in the POS (Fig. 6d); they are present in the support as well as the overgrowth. Na and Cl are always closely associated, suggesting the presence of halite (NaCl). They were detected along grain boundaries around a pore. In summary, the SEM–EDS analyses show there is no unequivocal difference in trace element content between the POS support and overgrowth and that the difference between the two parts of the POS is only structural.

U-series chronology

U-series ages were derived from five subsamples of sample MLp1 (Fig. 4). Detailed uranium–thorium data for sample MLp1 (Fig. 4) are provided in Table 1. The soda straw stalactite, which represents growth above the water table, has an age of 5948 ± 228 yr. The stalactite grew when sea level was lower.

The U-series data for the four POS subsamples at 1, 4, 6, and 8 mm away from the stalactite (see Fig. 4) are all of similar age, if considering their error (Table 1). We used these results to produce a 232Th/238U–234U/238U–230Th/238U isochron that yielded an age of 2795 ± 88 yr. IsoplotR offers a routine, which assumes that analytical uncertainties are not representative of the true uncertainties and applies an overdispersion term that reduces the mean square weighted deviation (MSWD) to unity (Vermeesch, Reference Vermeesch2018), with the interpretation that the larger uncertainties assigned by the program are due to geologic scatter. In our case, three of the four analyses produced an isochron age of 2852 ± 270 yr (MSWD = 9) without any assumptions. Applying the 4-point isochron-based detrital thorium correction, we use the weighted average age from the four U-series dates for the POS of 2759 ± 140 yr (Fig. 7).

Figure 7. The 232Th/238U–234U/238U–230Th/238U isochron age for the phreatic overgrowth of samples MLp1. (a) The over dispersion routine provided in IsoplotR (Vermeesch, Reference Vermeesch2018) produces an age and 2σ absolute uncertainty of 2794 ± 88 years. (b) The isochron produced from three of the four subsamples yields an age of 2852 ± 270 years.

DISCUSSION

POS formation

Morphology of the overgrowth depends on the tide-controlled daily groundwater level fluctuation, substrate shape, length of growth, and the degree of immersion of the substrate at the time of formation (Vesica et al., Reference Vesica, Tuccimei, Turi, Fornós, Ginés and Ginés2000). The development of phreatic overgrowth (MLp1) around a regularly shaped pre-existing vadose soda straw is linked to tidal fluctuations, as depicted in Figure 1. The MLp1 overgrowth (Figs. 4 and 6) grew over a pre-existing soda straw that was not long enough to capture the lowest tidal range. In such a case, POS deposition does not coincide with the full sea-level fluctuation range but likely records only the upper part of that range (Vesica et al., Reference Vesica, Tuccimei, Turi, Fornós, Ginés and Ginés2000; Ginés et al., Reference Ginés, Ginés, Fornós, Tuccimei, Onac, Gràcia, Ginés, Ginés, Gómez-Pujol, Onac and Fornós2012). Based on the morphology of the MLp1 overgrowth, it is apparent that the POS deposition records fluctuations between mean sea level (± 0.15 m) and the high tide, which is 24 cm above current mean sea level, given the average tidal range. We presumed that the asymmetric, almost flat-bottomed shape of the POS indicates its closeness to the mean sea level, as described in Tuccimei et al. (Reference Tuccimei, Soligo, Ginés, Ginés, Fornós, Kramers and Villa2010) and Ginés et al. (Reference Ginés, Ginés, Fornós, Tuccimei, Onac, Gràcia, Ginés, Ginés, Gómez-Pujol, Onac and Fornós2012).

XRD analysis of MLp1 shows that the fibrous overgrowth is made of calcite. Such overgrowths are also known from Mallorca, although the majority of the POS with fibrous crystals on this island are aragonite (Ginés et al., Reference Ginés, Ginés, Fornós, Tuccimei, Onac, Gràcia, Ginés, Ginés, Gómez-Pujol, Onac and Fornós2012). The calcite overgrowths on Mallorca, however, showed higher Mg concentrations than the calcitic support, which was attributed to the Mg-richer brackish water from which the overgrowth calcite precipitated (Vesica et al., Reference Vesica, Tuccimei, Turi, Fornós, Ginés and Ginés2000; Ginés et al., Reference Ginés, Fornós, Ginés, Sanjaume and Mateu2005; Csoma et al., Reference Csoma, Goldstein and Pomar2006). The absence of any distinct difference in trace element content between the support and the overgrowth in MLp1, as revealed by XRF and SEM-EDS analyses, therefore indicates that the drip water and pool water from which the calcite precipitated must have had a similar composition. In other words, the top-most layer of the cave pool water column was probably pure fresh groundwater unmixed with marine water. The often-observed epitaxial growth of the crystals in POS over the supporting soda straw (Fig. 6b) indicates a transitional phase in which the crystals were growing in alternating conditions (i.e., as support [while emerged during low tide] and overgrowth [when submerged during high tide] crystals). This may in turn explain the occurrence of the distinct dark overgrowth layer observed with the naked eye around the support (Fig. 4). This layer might have formed during this transitional period.

Time of deposition and regional relative sea-level context

Studies on Holocene relative sea-level changes in the northern Adriatic have a long history. This research began with the first evidence of submerged archaeological structures, as documented by Gnirs (Reference Gnirs1908) and Degrassi (Reference Degrassi1955). Subsequent investigations in Istria, including work by Antonioli et al. (Reference Antonioli, Anzidei, Lambeck, Auriemma, Gaddi, Furlani and Orru2007), Faivre et al. (Reference Faivre, Fouache, Ghilardi, Antonioli, Furlani and Kovačić2011a), and Florido et al. (Reference Florido, Auriemma, Faivre, Radić Rossi, Antonioli, Furlani and Spada2011), have provided detailed insights into the last 1600 yr of relative sea-level changes in this region (Faivre et al., Reference Faivre, Bakran-Petricioli, Barešić, Horvatić and Macario2019) (Fig. 8). Recent high-resolution relative sea-level studies have further enriched our understanding of this complex phenomenon (Faivre et al., Reference Faivre, Bakran-Petricioli, Barešić, Horvatić and Macario2019; Kaniewski et al., Reference Kaniewski, Marriner, Cheddadi, Morhange, Vacchi, Rovere and Faivre2021). New paleoenvironmental reconstructions are now available from Cres Island covering the Late Pleistocene and Holocene (Brunović et al., Reference Brunović, Miko, Ilijanić, Peh, Hasan, Kolar, Šparica Miko and Razum2019, Reference Brunović, Miko, Hasan, Papatheodorou, Ilijanić, Miserocchi, Correggiari and Geraga2020) and from the more distant island of Pag (Ilijanić et al., Reference Ilijanić, Miko, Ivkić Filipović, Hasan, Šparica Miko, Petrinec, Terzić and Marković2022). However, data during the period of POS formation reported here are sparse. Thus, POS can provide new evidence, which can supplement and improve existing data.

Figure 8. Phreatic overgrowth sample MLp1 and speleothem L-1 superimposed on the relative sea-level curve for the northern Adriatic Sea constructed for Istria (Faivre et al., Reference Faivre, Fouache, Ghilardi, Antonioli, Furlani and Kovačić2011a, 2019).

Previous studies of speleothem deposition in Medvjeđa Špilja, as well as in other Croatian coastal caves (e.g., Surić et al., Reference Surić, Jalžić and Petricioli2007; Surić and Juračić, Reference Surić and Juračić2010), revealed hiatuses in submerged stalagmites that can be used in studies of relative sea-level change. According to those records, ca. 7000 yr ago, sea level was ~10 m lower than present, while ca. 3350 yr ago sea level rose to around −1.5 m (Surić et al., Reference Surić, Jalžić and Petricioli2007). Speleothem L-1 of Surić et al. (Reference Surić, Jalžić and Petricioli2007) is a stalactite that had broken off from the roof and was found in an upside-down position. Detailed analysis revealed that calcite deposition continued in the new position, in the form of needle-like deposits, around a previously deposited stalactite, indicating alternating freshwater/brackish conditions. Consequently, Surić et al. (Reference Surić, Jalžić and Petricioli2007) and Surić and Juračić (Reference Surić and Juračić2010) proposed that the sea level ca. 3350 years cal yr BP had not yet reached −1.5 m.

The supporting soda straw in the MLp1 sample is 5948 ± 228 years old, indicating vadose conditions during its formation. The four ages of the POS are slightly reversed, but within their errors they are essentially the same. Brackish water has to remain stable for a length of time to become saturated with enough calcite for deposition that would be recognized as a POS (Polyak et al., Reference Polyak, Onac, Fornós, Hay, Asmerom, Dorale, Ginés, Tuccimei and Ginés2018). Therefore, we estimate that around 2759 ± 140 yr sea level was relatively stable for about 300 years at −1.28 m ± 0.15 m below current mean sea level. Based on our new findings of POS deposition, we also presume that the needle-like deposits of L-1 described by Surić and Juračić (Reference Surić and Juračić2010) represent POS that formed within the tidal range.

According to isotopic records of eastern Adriatic speleothems, the Holocene is characterized by many and sudden environmental changes, whereas the Late Holocene primarily is characterized by drier conditions (Surić et al., Reference Surić, Columbu, Lončarić, Bajo, Bočić, Lončar, Drysdale and Hellstrom2021), which were occasionally interrupted by wet stages (Lončar et al., Reference Lončar, Bar-Matthews, Ayalon, Surić and Faivre2017, 2019). Periods of relative sea-level stability have been documented along the eastern Adriatic coast using different RSL markers such as tidal notches (e.g., Fouache et al., Reference Fouache, Faivre, Dufaure, Kovačić and Tassaux2000; Antonioli et al., Reference Antonioli, Carulli, Furlani, Auriemma and Marocco2004b; Benac et al., Reference Benac, Juračić and Bakran-Petricioli2004; Marriner et al., Reference Marriner, Morhange, Faivre, Flaux, Vacchi, Miko, Boetto and Radić Rossi2014) and algal rims (e.g., Faivre et al., Reference Faivre, Bakran-Petricioli, Barešić, Horvatić and Macario2019, Reference Faivre, Bakran-Petricioli, Barešić and Horvatić2021a, Reference Faivre, Bakran-Petricioli, Herak, Barešić and Borkovićb). Eastern Adriatic tidal notches are interpreted to have formed during two main periods (Late Antique Little Ice Age and Little Ice Age) of relative sea-level stability, similar to findings in other parts in the eastern (e.g., Boulton and Stewart, Reference Boulton and Stewart2015), central (e.g., Faivre et al., Reference Faivre, Bakran-Petricioli, Horvatinčić and Sironić2013, Reference Faivre, Bakran-Petricioli, Herak, Barešić and Borković2021b), and western Mediterranean (e.g., Vacchi et al., Reference Vacchi, Gatti, Kulling, Morhange and Marriner2022). The northern Adriatic notches formed during the Late Antique Little Ice Age (Faivre et al., Reference Faivre, Bakran-Petricioli, Barešić, Horvatić and Macario2019), whereas central Adriatic notches formed during the Little Ice Age, about 500 years ago (Faivre et al., Reference Faivre, Bakran-Petricioli, Horvatinčić and Sironić2013, Faivre and Butorac, Reference Faivre and Butorac2018). These periods of relative sea-level stability have also been observed in the southern Adriatic (Faivre et al., Reference Faivre, Bakran-Petricioli, Barešić and Horvatić2021a, Reference Faivre, Bakran-Petricioli, Herak, Barešić and Borkovićb). Such intervals could be related to periods of drop in global mean sea level connected to the Northern Hemisphere global mean cooling noted by Mann et al. (Reference Mann, Zhang, Hughes, Bradley, Miller, Rutherford and Ni2008), Ljungqvist (Reference Ljungqvist2010), and PAGES 2k Consortium (2013, 2019), which offset the glacial isostatic adjustment effects (Faivre et al., Reference Faivre, Bakran-Petricioli, Kaniewski, Marriner, Tomljenović, Sečanj, Horvatić, Barešić, Morhange and Drysdale2023).

A period of relative sea-level stability during the Late Bronze Age and the transition to the Iron Age between 3250 and 2800 cal yr BP was already documented in the southern Adriatic based on the presence of algal rims on Lopud Island (Faivre et al., Reference Faivre, Bakran-Petricioli, Barešić and Horvatić2021a). Thus, POS data from the Medvjeđa Špilja (formed around 2.8 ka) provide possible indications of RSL stability in the northern Adriatic Sea at the end of this period. Dry conditions during that period were inferred from different proxies throughout the Mediterranean, including the SPD-1 stalagmite from the island of Dugi otok. SPD-1 shows that the entire period between ca. 3.3 and 2.7 ka was dry, although it was interrupted by short wet events, and true wetter conditions only followed after 2.7 ka (Lončar et al., Reference Lončar, Bar-Matthews, Ayalon, Faivre and Surić2019). Particularly prominent dry conditions around ca. 3300 cal yr BP were also observed in lake sediments from Albania and Montenegro (Zanchetta et al., Reference Zanchetta, Van Welden, Baneschi, Drysdale, Sadori, Roberts, Giardini, Beck, Pascucci and Sulpizio2012).

This dry period can also be associated with the cooling phase in the Aegean Sea, also around 3300 cal BP (Rohling et al., Reference Rohling, Mayewsky, Hayes, Abu-Zied and Casford2002), and with the severe long-term drought in the Eastern Mediterranean, which dramatically affected agriculture and triggered societal collapse in the Late Bronze and Iron ages, generally between 3150–2800 cal yr BP (Kaniewski et al., Reference Kaniewski, Paulissen, Van Campo, Weiss, Otto, Bretschneider and Van Lerberghe2010; Kagan et al., Reference Kagan, Langgut, Boaretto, Neumann and Stein2015; Langgut et al., Reference Langgut, Finkelstein, Litt, Neumann and Stein2015). Overall, the formation of our POS could be roughly related to the end of the longest Holocene cooling phase in the Mediterranean associated with the 3.2-ka event characterized by cooling of −0.38 ± 0.19°C, over ca. 320 years (Marriner et al., Reference Marriner, Kaniewski, Pourkerman and Devillers2022).

The above studies directly indicate that relative sea-level change during the Late Holocene was not linear and provide evidence for the existence of periods of relative sea-level stability that are likely related to climate conditions (Faivre et al., Reference Faivre, Bakran-Petricioli, Kaniewski, Marriner, Tomljenović, Sečanj, Horvatić, Barešić, Morhange and Drysdale2023). Overall, our new results from Mali Lošinj Island POS correspond well to the nearby 5000-yr composite relative sea-level curve from Istria (Faivre et al., Reference Faivre, Fouache, Ghilardi, Antonioli, Furlani and Kovačić2011a, 2019; Kaniewski et al., Reference Kaniewski, Marriner, Cheddadi, Morhange, Vacchi, Rovere and Faivre2021), suggesting a similar trend of subsidence in Istria and along the eastern coast of Lošinj during last 2800 years of ~0.46 mm/yr (Fig. 8). The general agreement of the results obtained from POS with previous local and regional results from the Adriatic and the eastern Mediterranean confirm POS as a reliable sea-level indicator.

CONCLUSIONS

Speleothem-based research into Pleistocene and Holocene relative sea-level changes along the eastern Adriatic has traditionally focused on biogenic encrustations, identified hiatuses, and mineralogical shifts within submerged speleothems. Our research in the Medvjeđa Špilja (Lošinj Island, northern Adriatic) reveals that the cave hosts a pool environment favorable for phreatic overgrowths on speleothems (POS) precipitation during sea-level stillstand conditions. Mineralogical analysis confirmed that speleothem MLp-1 is made up of calcite and has a typical morphology of POS, which makes it the first POS found in the Adriatic. The XRF and SEM–EDS analyses showed that there is no difference in trace element composition between the support and the overgrowth. Obtained results suggests that the drip water from which the support (calcite soda straw) formed and the upper part of the water column in the cave pool (from which the overgrowth precipitated) had the same chemical composition. Based on uranium-series dating, we conclude that the relative sea level at ca. 2.8 ka must have remained stable for ca. 300 years at a depth of approximately −1.28 ± 0.15 m below the current MSL. Patterns of relative sea-level changes along the eastern coast of Lošinj Island align with trends seen along the Istrian coast. This suggests a general subsidence rate of ~0.46 mm/yr during the Late Holocene in the study area. Further research of POS at this study site will enable recording more sea-level index points, which will enable the development of longer and more precise curves of relative sea-level change in the northern Adriatic.

Acknowledgments

The research was conducted within the SEALeveL project (HRZZ-IP- 2019-04-9445) funded by Croatian Science Foundation. B.P.O. and V.J.P. were funded by a collaborative NSF grant (AGS 2202683 and 2202712). We are grateful to cave divers Petra Kovač-Konrad and Marko Baričević for cave prospecting and sampling assistance; safety cave diver Vladan Strigo; cavers from the Rijeka and Lošinj Croatian Mountain Rescue Team: Dino Groznić, Marko Župančić, David Mitrović, and Luka Petrinić; Anel Hasić, Zrinka Ettinger Starčić, Neno Starčić (Diving Center Sub Season), Mr. Velimir Milenović and his family for field assistance, and Ivica Rendulić for the help with illustrations. We would like to thank the Senior Editor Derek Booth, Guest Editor Andrea Columbu, and the reviewers (Rieneke Weij and one anonymous) whose inputs have significantly improved our manuscript.

References

REFERENCES

Alessio, M., Allegri, L., Antonioli, F., Belluomini, G., Ferranti, L., Importa, S., Manfra, L., Proposito, A., 1992. Risultati preliminari relativi alla datazione di speleotemi sommersi nelle fasce costiere del Tirreno centrale. Giornale di Geologia ser. 3, 54, 165193.Google Scholar
Alessio, M., Allegri, L., Antonioli, F., Belluomini, G., Improta, S., Manfra, L., Preite Martinez, M., 1994. La curva di risalita del mare Tirreno negli ultimi 43 ka ricavata da datazioni su speleotemi sommersi e dati archeologici. Memorie Descrittive Della Carta Geologica d'Italia LII, 261276.Google Scholar
Antonioli, F., Anzidei, M., Lambeck, K., Auriemma, R., Gaddi, D., Furlani, S., Orru, P., et al., 2007. Sea level change during the Holocene in Sardinia and in the northeastern Adriatic (central Mediterranean Sea) from archaeological and geomorphological data. Quaternary Science Reviews 26, 24632486.Google Scholar
Antonioli, F., Bard, E., Potter, E.-K., Silenzi, S., Improta, S., 2004a. 215-ka history of sea-level oscillations from marine and continental layers in Argentarola Cave speleothems (Italy). Global and Planetary Change 43, 5778.Google Scholar
Antonioli, F., Carulli, G.B., Furlani, S., Auriemma, R., Marocco, R., 2004b. The enigma of the submerged marine notches in the northern Adriatic Sea. Quaternaria Nova 8, 263275.Google Scholar
Antonioli, F., Cremona, G., Immordino, F., Puglisi, C., Romagnoli, C., Silenzi, S., Valpreda, E., Verrubbi, V., 2002. New data on the Holocenic sea-level rise in NW Sicily (central Mediterranean Sea). Global and Planetary Change 34, 121140.Google Scholar
Antonioli, F., Furlani, S., Montagna, P., Stocchi, P., 2021. The use of submerged speleothems for sea level studies in the Mediterranean Sea: a new perspective using glacial isostatic adjustment (GIA). Geosciences 11, 77. https://doi.org/10.3390/geosciences11020077.Google Scholar
Antonioli, F., Oliverio, M., 1996. Holocene sea-level rise recorded by a radiocarbon-dated mussel in a submerged speleothem beneath the Mediterranean Sea. Quaternary Research 45, 241244.Google Scholar
Asmerom, Y., Polyak, V., Schwieters, J., Bouman, C., 2006. Routine high-precision U–Th isotope analyses for paleoclimate chronology. Geochemica et Cosmochimica Acta 70, A24. https://doi.org/10.1016/j.gca.2006.06.061.Google Scholar
Bard, E., Antonioli, F., Silenzi, S., 2002. Sea-level during the penultimate interglacial period based on a submerged stalagmite from Argentarola Cave (Italy). Earth and Planetary Science Letters 196, 135146.Google Scholar
Benac, Č., Juračić, M., Bakran-Petricioli, T., 2004. Submerged tidal notches in the Rijeka Bay NE Adriatic Sea: indicators of relative sea-level change and of recent tectonic movements. Marine Geology 212, 2133.Google Scholar
Bonacci, O., Roje-Bonacci, T., 2003. Groundwater on small Adriatic islands. RMZ – Materials and Geoenvironment 50, 4144.Google Scholar
Boulton, S.J., Stewart, I.S., 2015. Holocene coastal notches in the Mediterranean region: indicators of palaeoseismic clustering? Geomorphology 237, 2937.Google Scholar
Brunović, D., Miko, S., Hasan, O., Papatheodorou, G., Ilijanić, N., Miserocchi, S., Correggiari, A., Geraga, M., 2020. Late Pleistocene and Holocene paleoenvironmental reconstruction of a drowned karst isolation basin (Lošinj Channel, NE Adriatic Sea). Palaeogeography, Palaeoclimatology, Palaeoecology 544, 109587. https://doi.org/10.1016/j.palaeo.2020.109587.Google Scholar
Brunović, D., Miko, S., Ilijanić, N., Peh, Z., Hasan, O., Kolar, T., Šparica Miko, M., Razum, I., 2019. Holocene foraminiferal and geochemical records in the coastal karst dolines of Cres Island, Croatia. Geologia Croatica 72, 1942.Google Scholar
Cheng, H., Edwards, R.L., Shen, C.C., Polyak, V.J., Asmerom, Y., Woodhead, J., Hellstrome, J., et al., 2013. Improvements in 230Th dating, 230Th and 234U half-life values, and U–Th isotopic measurements by multi-collector inductively coupled plasma mass spectrometry. Earth and Planetary Science Letters 371, 8291.Google Scholar
CMHS, 2022. Croatian Meteorological and Hydrological Service. https://meteo.hr/index_en.php.Google Scholar
Csoma, A.É., Goldstein, R.H., Pomar, L., 2006. Pleistocene speleothems of Mallorca: implications for palaeoclimate and carbonate diagenesis in mixing zones. Sedimentology 53, 213236.Google Scholar
Degrassi, A., 1955. I Porti Romani Dell'Istria. Anthemon, Firenze, 169 pp.Google Scholar
de Joly, R., 1929. Explorations spéléologiques à Majorque (1929). Revue de Géographie Physique et de Géologie Dynamique 2, 233245.Google Scholar
De Waele, J., D'Angeli, I.M., Bontognali, T., Tuccimei, P., Scholz, D., Jochum, K.P., Columbu, A., et al., 2018. Speleothems in a north Cuban cave register sea-level changes and Pleistocene uplift rates. Earth Surface Processes and Landforms 43, 23132326.Google Scholar
De Waele, J., D'Angeli, I.M., Tisato, N., Tuccimei, P., Soligo, M., Ginés, J., Ginés, A., et al., 2017. Coastal uplift rate at Matanzas (Cuba) inferred from MIS5e phreatic overgrowths on speleothems. Terra Nova 29, 98105.Google Scholar
Dorale, J.A., Onac, B.P., Fornós, J.J., Ginés, J., Ginés, A., Tuccimei, P., Peate, D.W., 2010. Sea-level highstand 81,000 years ago in Mallorca. Science 327, 860863.Google Scholar
Dumitru, O.A., Polyak, V.J., Asmerom, Y., Onac, B.P., 2021. Last interglacial sea-level history from speleothems: a global standardized database. Earth System Science Data 13, 20772094.Google Scholar
Faivre, S., Bakran-Petricioli, , Kaniewski, D., Marriner, N., Tomljenović, B., Sečanj, M., Horvatić, D., Barešić, J., Morhange, C., Drysdale, R.N., 2023. Driving processes of relative sea-level change in the Adriatic during the past two millennia: from local tectonic movements in the Dubrovnik archipelago (Jakljan and Šipan islands) to global mean sea level contributions (Central Mediterranean). Global and Planetary Change 227, 104158.Google Scholar
Faivre, S., Bakran-Petricioli, T., Barešić, J., Horvatić, D., 2021a. Lithophyllum rims as biological markers for constraining palaeoseismic events and relative sea-level variations during the last 3.3 ka on Lopud Island, southern Adriatic, Croatia. Global and Planetary Change 202, 103517.Google Scholar
Faivre, S., Bakran-Petricioli, T., Barešić, J., Horvatić, D., Macario, K., 2019. Relative sea-level change and climate change in the northeastern Adriatic during the last 1.5 ka (Istria, Croatia). Quaternary Science Reviews 222, 105909.Google Scholar
Faivre, S., Bakran-Petricioli, T., Herak, M., Barešić, J., Borković, D., 2021b. Late Holocene interplay between coseismic uplift events and interseismic subsidence at Koločep Island and Grebeni islets in the Dubrovnik archipelago (southern Adriatic, Croatia). Quaternary Science Reviews 274, 107284.Google Scholar
Faivre, S., Bakran-Petricioli, T., Horvatinčić, N., Sironić, A., 2013. Distinct phases of relative sea level changes in the central Adriatic during the last 1500 years—influence of climatic variations? Palaeogeography, Palaeoclimatology, Palaeoecology 369, 163174.Google Scholar
Faivre, S., Butorac, V., 2018. Recently submerged tidal notches in the wider Makarska area (central Adriatic, Croatia). Quaternary International 494, 225235.Google Scholar
Faivre, S., Fouache, E., 2003. Some tectonic influences on the Croatian shoreline evolution in the last 2000 years. Zeitschrift für Geomorphologie 47, 521537.Google Scholar
Faivre, S., Fouache, E., Ghilardi, M., Antonioli, F., Furlani, S., Kovačić, V., 2011a. Relative sea level change in western Istria (Croatia) during the last millennium. Quaternary International 232, 132143.Google Scholar
Faivre, S., Pahernik, M., Maradin, M., 2011b. The gully of Potovošća on the Island of Krk – the effects of a short-term event. Geologia Croatica 64, 6476.Google Scholar
Florido, E., Auriemma, R., Faivre, S., Radić Rossi, I., Antonioli, F., Furlani, S., Spada, G., 2011. Istrian and Dalmatian fishtanks as sea level markers. Quaternary International 232, 105113.Google Scholar
Fornós, J. J., Gelabert, B., Ginés, A., Ginés, J., Tuccimei, P., Vesica, P., 2002. Phreatic overgrowths on speleothems: a useful tool in structural geology in littoral karstic landscapes. The example of eastern Mallorca (Balearic Islands). Geodinamica Acta 15, 113125.Google Scholar
Fouache, E., Faivre, S., Dufaure, J-J., Kovačić, V., Tassaux, F., 2000. New observations on the evolution of the Croatian shoreline between Poreč and Zadar over the past 2000 years. Zeitschrift für Geomorphologie 122 (supplement), 346.Google Scholar
Gascoyne, M., Benjamin, G.J., Schwarcz, H.P., Ford, D.C., 1979. Sea-level lowering during the Illinoian Glaciation: evidence from a Bahama “Blue Hole”. Science 205, 806808.Google Scholar
Ginés, A., Ginés, J., 1974. Consideraciones sobre los mecanismos de fosilización de la Cova de sa Bassa Blanca y su paralelismo con las formaciones marinas del Cuaternario. Boletin de la Sociedad de Historia Natural de las Baleares 19, 1128.Google Scholar
Ginés, J., Fornós, J.J., Ginés, Á., 2005. Els espeleotemes freàtics del Quaternari de Mallorca: aspectes morfològics, mineralògics i cristalogràfics. In: Sanjaume, E., Mateu, J.F. (Eds.), Geomorfologia Litoral i Quaternari. Homenatge al Professor Vicenç M. Rosselló i Verger. Universitat de València, València, pp. 151165. [in Catalan]Google Scholar
Ginés, J., Ginés, A., Fornós, J.J., Tuccimei, P., Onac, B.P., Gràcia, F., 2012. Phreatic overgrowths on speleothems (POS) from Mallorca, Spain: updating forty years of research. In: Ginés, A., Ginés, J., Gómez-Pujol, L., Onac, B.P., Fornós, J.J. (Eds.), Mallorca: A Mediterranean Benchmark for Quaternary Studies. Monografies de la Societat d'Història Natural de les Balears 18, 111146.Google Scholar
Ginés, J., Ginés, A., Pomar, L., 1981. Morphological and mineralogical features of phreatic speleothems occurring in coastal caves of Majorca (Spain). In: Beck, B.F. (Ed.), Proceedings of the Eighth International Congress of Speleology, Bowling Green, Kentucky 2, 529532.Google Scholar
Gnirs, A., 1908. Beobachtungen über den Fortschritt einer säkularen Niveauschwankung des Meeres während des letzten zwei Jahrtausende. Mitteilungen der Kaiserlich–Königlichen Geographischen Gesellschaft 51, 156.Google Scholar
Harmon, R.S., Schwarcz, H.P., Ford, D.C., 1978. Late Pleistocene sea level history of Bermuda. Quaternary Research 9, 205218.Google Scholar
Herwegh, M., 2000. A new technique to automatically quantify microstructures of fine grained carbonate mylonites: two-step etching combined with SEM imaging and image analysis. Journal of Structural Geology 22, 391400.Google Scholar
Ilijanić, N., Miko, S., Ivkić Filipović, I., Hasan, O., Šparica Miko, M., Petrinec, B., Terzić, J., Marković, T., 2022. A Holocene sedimentary record and the impact of sea-level rise in the karst lake Velo Blato and the wetlands on Pag Island (Croatia). Water 14, 342. https://doi.org/10.3390/w14030342.Google Scholar
Jalžić, B., 2007. Medvjeđa Špilja na otoku Lošinju. Speleolog 55, 4555. [in Croatian]Google Scholar
Jenson, A., Schwartz, B., Li, Y., Gao, Y., 2018. The implications and limitations of phreatic overgrowths of speleothems as sea level indicators: Quintana Roo, Mexico. Geological Society of America Abstracts with Programs 50 (6), 147–146. https://doi.org/10.1130/abs/2018AM-318501.Google Scholar
Kagan, E.J., Langgut, D., Boaretto, E., Neumann, F.H., Stein, M., 2015. Dead Sea levels during the Bronze and Iron ages. Radiocarbon 57, 237252.Google Scholar
Kaniewski, D., Marriner, N., Cheddadi, R., Morhange, C., Vacchi, M., Rovere, A., Faivre, S., et al., 2021. Coastal submersions in the north-eastern Adriatic during the last 5200 years. Global and Planetary Change 204, 103570. doi.org/10.1016/j.gloplacha.2021.103570.Google Scholar
Kaniewski, D., Paulissen, E., Van Campo, E., Weiss, H., Otto, T., Bretschneider, J., Van Lerberghe, K., 2010. Late second–early first millennium BC abrupt climate changes in coastal Syria and their possible significance for the history of the Eastern Mediterranean. Quaternary Research 74, 207215.Google Scholar
Korbar, T., 2009. Orogenic evolution of the External Dinarides in the NE Adriatic region: a model constrained by tectonostratigraphy of Upper Cretaceous to Paleogene carbonates. Earth-Science Reviews 96, 296312.Google Scholar
Langgut, D, Finkelstein, I., Litt, T., Neumann, F.H., Stein, M., 2015. Vegetation and climate changes during the Bronze and Iron ages (~3600–600 BCE) in the southern Levant based on palynological records. Radiocarbon 57, 217235.Google Scholar
Li, W.X., Lundberg, J., Dickin, A.P., Ford, D.C., Schwarcz, H.P., McNutt, R., Williams, D., 1989. High-precision mass-spectrometric uranium-series dating of cave deposits and implications for palaeoclimate studies. Nature 339, 534536.Google Scholar
Ljungqvist, F.C., 2010. A new reconstruction of temperature variability in the extratropical Northern Hemisphere during the last two millennia. Geografiska Annaler 92, 339351.Google Scholar
Lončar, N., Bar-Matthews, M., Ayalon, A., Faivre, S., Surić, M., 2019. Holocene climatic conditions in the eastern Adriatic recorded in stalagmites from Strašna Peć Cave (Croatia). Quaternary International 508, 98106.Google Scholar
Lončar, N., Bar-Matthews, M., Ayalon, A., Surić, M., Faivre, S., 2017. Early and mid-Holocene environmental conditions in the eastern Adriatic recorded in speleothems from Mala Špilja Cave and Velika Špilja cave (Mljet Island, Croatia). Acta Carsologica 46, 229249.Google Scholar
Malez, M., Božičević, S., 1965. The Medvjeđa Pećina (Bear Cave) on Lošinj Island, a rare case of submerged cave. In: Stelcl, O. (Ed.), International Speleological Conference, Brno, Problems of Speleological Research, 29th June–4th July 1964, Prague, pp. 211216.Google Scholar
Malez, M., Sliepčević, A., Srdoč, D., 1979. Određivanje starosti metodom radioaktivnog ugljika kvartarnim naslagama na nekim lokalitetima u Dinarskom kršu [Radiocarbon dating of Quaternary deposits on some localities in Dinaric karst]. Rad JAZU, 383, Razred za prirodne znanosti 18, 227271. [in Croatian]Google Scholar
Mann, M.E., Zhang, Z., Hughes, M.K., Bradley, R.S., Miller, S.K., Rutherford, S., Ni, F., 2008. Proxy-based reconstructions of hemispheric and global surface temperature variations over the past two millennia. Proceedings of the National Academy of Sciences USA 105, 1325213257.Google Scholar
Marriner, N., Kaniewski, D., Pourkerman, M., Devillers, B., 2022. Anthropocene tipping point reverses long-term Holocene cooling of the Mediterranean Sea: a meta-analysis of the basin's sea surface temperature records. Earth-Science Reviews 227, 103986. https://doi.org/10.1016/j.earscirev.2022.103986.Google Scholar
Marriner, N., Morhange, C., Faivre, S., Flaux, C., Vacchi, M., Miko, S., Boetto, G., Radić Rossi, I., 2014. Post-Roman sea-level changes on Pag Island (Adriatic Sea): dating Croatia's “enigmatic” coastal notch? Geomorphology 221, 8394.Google Scholar
Miklavič, B., Yokoyama, Y., Urata, K., Miyairi, Y., Kan, H., 2018. Holocene relative sea level history from phreatic overgrowths on speleothems (POS) on Minami Daito Island, Northern Philippine Sea. Quaternary International 471, 359368.Google Scholar
Onac, B.P., Ginés, A., Ginés, J., Fornós, J.J., Dorale, J.A., 2012. Late Quaternary sea-level history: a speleothem perspective. In Ginés, A., Ginés, J., Gómez-Pujol, L., Onac, B.P., Fornós, J.J. (Eds.), Mallorca: A Mediterranean Benchmark for Quaternary Studies. Monografies de la Societat d'Història Natural de les Balears 18, 147161.Google Scholar
Onac, B.P., Mitrovica, J.X., Ginés, J., Asmerom, Y., Polyak, V.J., Tuccimei, P., Ashe, E.L., et al., 2022. Exceptionally stable preindustrial sea level inferred from the western Mediterranean Sea. Science Advances 8, eabm6185. https://doi.org/10.1126/sciadv.abm6185.Google Scholar
Onset, 2022., HOBO® U20 Water Level Logger (U20-001-0x and U20-001-0x-Ti) Manual. https://www.onsetcomp.com/sites/default/files/resources-documents/12315-J%20U20%20Manual.pdf.Google Scholar
PAGES 2k Consortium, 2013. Continental-scale temperature variability during the past two millennia. Nature Geoscience 6, 339346.Google Scholar
PAGES 2k Consortium, 2019. Consistent multidecadal variability in global temperature reconstructions and simulations over the Common Era. Nature Geoscience 12, 643649.Google Scholar
Polyak, V.J., Onac, B.P., Fornós, J.J., Hay, C., Asmerom, Y., Dorale, JA., Ginés, J., Tuccimei, P., Ginés, A., 2018. A highly resolved record of relative sea level in the western Mediterranean Sea during the last interglacial period. Nature Geoscience 11, 860864.Google Scholar
Rodés, L., 1925. Los cambios de nivel en las Cuevas del Drach (Manacor, Mallorca) y suoscilación rítmica de 40 minutos. Memorias de la Real Academia de Ciencias y Artes de Barcelona 19, 207221.Google Scholar
Rohling, E.J., Mayewsky, P.A., Hayes, A., Abu-Zied, R.H., Casford, J.S.L., 2002. Holocene atmosphere–ocean interactions: records from Greenland and the Aegean Sea. Climate Dynamics 18, 587593.Google Scholar
Schmid, S.M., Fügenschuh, B., Kounov, A., Matenco, L., Nievergelt, P., Oberhänsli, R., Pleuger, J., et al., 2020. Tectonic units of the Alpine collision zone between Eastern Alps and western Turkey. Gondwana Research 78, 308374.Google Scholar
Shennan, I., Long, A., Benjamin, A., Horton, P., 2015. Introduction. In: Shennan, I., Long, A., Benjamin, A., Horton, P. (Eds.), Handbook of Sea-Level Research. John Wiley & Sons, Hoboken, pp. 12.Google Scholar
Špelić, M., Del Ben, A., Petrinjak, K., 2021. Structural setting and geodynamics of the Kvarner area (northern Adriatic). Marine and Petroleum Geology 125, 104857. https://doi.org/10.1016/j.marpetgeo.2020.104857.Google Scholar
Stocchi, P., Antonioli, F., Montagna, P., Pepe, F., Lo Presti, V., Caruso, A., Corradino, M., et al., 2017. A stalactite record of four relative sea-level highstands during the Middle Pleistocene Transition. Quaternary Science Reviews 173, 92100.Google Scholar
Surić, M., Columbu, A., Lončarić, R., Bajo, P., Bočić, N., Lončar, N., Drysdale, R., Hellstrom, J.C., 2021 - Holocene hydroclimate changes in continental Croatia recorded in speleothem δ13C and δ18O from Nova Grgosova Cave. The Holocene, 31, 14011416.Google Scholar
Surić, M., Jalžić, B., Petricioli, D., 2007. Submerged speleothems – expect the unexpected. Examples from the eastern Adriatic coast (Croatia). Acta Carsologica 36, 389396.Google Scholar
Surić, M., Juračić, M., 2010. Late Pleistocene–Holocene environmental changes – records from submerged speleothems along the eastern Adriatic coast (Croatia). Geologia Croatica 63, 155169.Google Scholar
Surić, M., Juračić, M., Horvatinčić, N., Krajcar Bronić, I., 2005. Late Pleistocene–Holocene sea-level rise and the pattern of coastal karst inundation: records from submerged speleothems along the eastern Adriatic coast (Croatia). Marine Geology 214, 163175.Google Scholar
Surić, M., Korbar, T., Juračić, M., 2014. Tectonic constraints on the Late Pleistocene–Holocene relative sea-level change along the north-eastern Adriatic coast (Croatia). Geomorphology 220, 93103.Google Scholar
Surić, M., Lončarić, R., Lončar, N., 2010. Submerged caves of Croatia: distribution, classification and origin. Environmental Earth Sciences 61, 14731480.Google Scholar
Surić, M., Richards, D.A., Hoffmann, D.L., Tibljaš, D., Juračić, M., 2009. Sea-level change during MIS 5a based on submerged speleothems from the eastern Adriatic Sea (Croatia). Marine Geology 262, 6267.Google Scholar
Tuccimei, P., Ginés, J., Delitala, M.C., Ginés, A., Gràcia, F., Fornós, J.J., Taddeucci, A., 2006. Last interglacial sea level changes in Mallorca Island (Western Mediterranean). High precision U-series data from phreatic overgrowths on speleothems. Zeitschrift für Geomorphologie 50, 121.Google Scholar
Tuccimei, P., Onac, B.P., Dorale, J.A., Ginés, J., Fornós, J.J., Ginés, A., Spada, G., Ruggieri, G., Mucedda, M., 2012. Decoding last interglacial sea-level variations in the western Mediterranean using speleothem encrustations from coastal caves in Mallorca and Sardinia: a field data–model comparison. Quaternary International 262, 5664.Google Scholar
Tuccimei, P., Soligo, M., Ginés, J., Ginés, A., Fornós, J., Kramers, J., Villa, I.M., 2010. Constraining Holocene sea levels using U–Th ages of phreatic overgrowths on speleothems from coastal caves in Mallorca (Western Mediterranean). Earth Surface Processes and Landforms 35, 782790.Google Scholar
Tuccimei, P., Van Strydonck, M., Ginés, A., Ginés, J., Soligo, M., Villa, I., Fornós, J., 2011. Comparison of 14C and U–Th ages of two Holocene phreatic overgrowths on speleothems from Mallorca (western Mediterranean): environmental implications. International Journal of Speleology 40, 18.Google Scholar
Vacchi, M., Gatti, G., Kulling, B., Morhange, C., Marriner, N., 2022. Climatic control on the formation of marine-notches in microtidal settings: new data from the northwestern Mediterranean Sea. Marine Geology 453, 106929. https://doi.org/10.1016/j.margeo.2022.106929.Google Scholar
van de Plassche, O. (Ed.). 1986. Sea-Level Research: A Manual for the Collection and Evaluation of Data. Geo Books, Norwich, UK, 618 pp.Google Scholar
van Hengstum, P.J., Richards, D.A., Onac, B.P., Dorale, J.A., 2015. Coastal caves and sinkholes. In: Shennan, I., Long, A.J., Horton, B.P. (Eds.), Handbook of Sea-Level Research. John Wiley & Sons, Hoboken, pp. 83103.Google Scholar
van Hinsbergen, D.J.J., Torsvik, T.H., Schmid, S.M., Matenco, L.C., Maffione, M., Vissers, R. Gürer, D., Spakman, W., 2020. Orogenic architecture of the Mediterranean region and kinematic reconstruction of its tectonic evolution since the Triassic. Gondwana Research 81, 79229.Google Scholar
Vermeesch, P., 2018. IsoplotR: a free and open toolbox for geochronology. Geoscience Frontiers 9, 14791493.Google Scholar
Vesica, P. L., Tuccimei, P., Turi, B., Fornós, J. J., Ginés, A., Ginés, J., 2000. Late Pleistocene paleoclimates and sea-level change in the Mediterranean as inferred from stable isotope and U-series studies of overgrowths on speleothems, Mallorca, Spain. Quaternary Science Reviews 19, 865879.Google Scholar
Vlahović, I., Tišljar, J., Velić, I., Matičec, D., 2005. Evolution of the Adriatic carbonate platform: palaeogeography, main events and depositional dynamics. Palaeogeography, Palaeoclimatology, Palaeoecology 220, 333360.Google Scholar
Zanchetta, G., Van Welden, A., Baneschi, I., Drysdale, R., Sadori, L., Roberts, N., Giardini, M., Beck, C., Pascucci, V., Sulpizio, R., 2012. Multiproxy record for the last 4500 years from Lake Shkodra (Albania/Montenegro). Journal of Quaternary Science 27, 780789.Google Scholar
Figure 0

Figure 1. Conceptual model of phreatic overgrowths on speleothem (POS) formation (after Dumitru et al., 2021); msl = mean sea level.

Figure 1

Figure 2. Study site and geologic setting (based on the Geological Map of the Republic of Croatia, scale 1:300 000, Croatian Geological Survey); MNE = Montenegro.

Figure 2

Figure 3. Medvjeđa Špilja (Bear Cave) longitudinal profile. (a, b) Cross-sectional and plan maps, (c) position in relation to the sea.

Figure 3

Figure 4. (a, b) External morphology and longitudinal section of phreatic overgrowth on speleothems (POS) sample MLp1, showing locations of samples for U-series dating (black marks); note the ~1-mm thick overgrowth layer immediately around the support (a pre-existing stalactite). (c) XRF spot analyses (red) on the halved and outer surface of the POS.

Figure 4

Table 1. Summary of the U-series measurement results for the phreatic overgrowth of speleothems sample MLp1.

Figure 5

Figure 5. Trace element content in the analyzed spots indicated in Figure 4 (red circles).

Figure 6

Figure 6. (a) The three parts of the POS as seen macroscopically (left) and under SEM (rigth; the area is indicated by white rectangle on the macroscopic picture). (b) The boundary between the support and the overgrowth with visible epitaxial crystal overgrowth in the inset picture. (c). SEM image showing the boundary (red dashed line) between the support and the dark overgrowth layer, and the EDS spot analysis at the support–overgrowth boundary (1 mm dark layer), showing Mg content. Observe the crystal zoning near spot #3. (d) EDS elemental map of the area shown in (a) showing a uniform distribution of Mg in the calcite across all three layers, and localized concentration of Si and Al.

Figure 7

Figure 7. The 232Th/238U–234U/238U–230Th/238U isochron age for the phreatic overgrowth of samples MLp1. (a) The over dispersion routine provided in IsoplotR (Vermeesch, 2018) produces an age and 2σ absolute uncertainty of 2794 ± 88 years. (b) The isochron produced from three of the four subsamples yields an age of 2852 ± 270 years.

Figure 8

Figure 8. Phreatic overgrowth sample MLp1 and speleothem L-1 superimposed on the relative sea-level curve for the northern Adriatic Sea constructed for Istria (Faivre et al., 2011a, 2019).