Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-28T21:36:16.331Z Has data issue: false hasContentIssue false

Exploring a Mallorca cave flooding during the Little Ice Age using nondestructive techniques on a stalagmite: micro-CT and XRF core scanning

Published online by Cambridge University Press:  17 October 2023

Mercè Cisneros*
Affiliation:
GRC Geociències Marines, Departament de Dinàmica de la Terra i de l'Oceà, Facultat de Ciències de la Terra, Universitat de Barcelona, c/ Martí i Franqués s/n, 08028 Barcelona, Spain Centre en Canvi Climàtic, Departament de Geografia, Facultat de Turisme i Geografia, Universitat Rovira i Virgili, c/ Joanot Martorell 15, 43480, Vila-seca, Tarragona, Spain
Isabel Cacho
Affiliation:
GRC Geociències Marines, Departament de Dinàmica de la Terra i de l'Oceà, Facultat de Ciències de la Terra, Universitat de Barcelona, c/ Martí i Franqués s/n, 08028 Barcelona, Spain
Jaime Frigola
Affiliation:
GRC Geociències Marines, Departament de Dinàmica de la Terra i de l'Oceà, Facultat de Ciències de la Terra, Universitat de Barcelona, c/ Martí i Franqués s/n, 08028 Barcelona, Spain
Ana Moreno
Affiliation:
Departamento de Procesos Geoambientales y Cambio Global, Instituto Pirenaico de Ecología–CSIC, Av. Montañana 1005, 50059, Zaragoza, Spain
Heather Stoll
Affiliation:
Departamento de Geología, Universidad de Oviedo, c/ Jesús Arias de Velasco s/n, 33005, Oviedo, Spain; and Department of Earth Sciences, ETH Zürich, Clausiusstrasse 25, 8092, Zurich, Switzerland
Joan J. Fornós
Affiliation:
Grup de Recerca en Ciències de la Terra, Universitat de les Illes Balears, ctra/ de Valldemossa, 07122, Mallorca, Spain
Javier Sigró
Affiliation:
Centre en Canvi Climàtic, Departament de Geografia, Facultat de Turisme i Geografia, Universitat Rovira i Virgili, c/ Joanot Martorell 15, 43480, Vila-seca, Tarragona, Spain
Mariano Barriendos
Affiliation:
IDAEA, Instituto de Diagnóstico Ambiental y Estudios del Agua, Consejo Superior de Investigaciones Científicas, c/ Jordi Girona 18-26, 08034 Barcelona, Spain
*
Corresponding author: Mercè Cisneros; Email: mercecisnerosb@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

This study focuses on characterizing a discontinuity within the Seán stalagmite (4.75–7.75 cm) by means of two nondestructive techniques: (1) high-resolution micro-computed tomography (micro-CT) and (2) X-ray fluorescence (XRF) core scanning (XRFCS). Micro-CT was used to study the stalagmite density, and XRFCS was applied to obtain the qualitative elemental composition and colour measurements. The new data obtained from nondestructive techniques have been combined with previously published geochemical data and fabric determinations from the same stalagmite found in Sa Balma des Quartó cave in Mallorca. The two methodologies applied in the present study have improved the characterization of the distinctive horizon. The micro-CT images identified the layer as a minor event due the high air content. The distinctive horizon is characterized by a high Ti-content, indicating the arrival of terrigenous particles. Based on those observations, together with the fact that the micrite layer appears filling the gaps between the older columnar fabric, we argue that the micrite layer may represent a major flooding event inside the cave after the year 1616 ± 23 CE and before the year 1623 ± 28 CE, which can be related to an extreme rainfall event. This hypothesis is further supported by the observed cave flooding during the autumn of 2018.

Type
Thematic Set: Speleothem Paleoclimate
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re- use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press on behalf of Quaternary Research Center

INTRODUCTION

The Mediterranean region is considered to be highly vulnerable to global warming (Giorgi., Reference Giorgi2006). It is considered a hot spot in terms of future climate change scenarios outlined by climate models, with expected drastic hydrological changes, such as intense precipitation events and extreme periods of drought (Gibelin and Deque, Reference Gibelin and Deque2003; Giorgi, Reference Giorgi2006; Ulbrich et al., Reference Ulbrich, May, Li, Lionello, Pinto, Somot, Lionello, Malanotte-Rizzoli and Boscolo2006; Giorgi and Lionello, Reference Giorgi and Lionello2008; Sheffield and Wood, Reference Sheffield and Wood2008; Mariotti et al., Reference Mariotti, Zeng, Yoon, Artale, Navarra, Alpert and Li2008; Drobinsky et al., Reference Drobinsky, Da Silva, Bastin, Mailler, Muller, Ahrens, Christensen and Lionello2020). Although the total annual precipitation in the Balearic region is decreasing, the contribution of minimum and maximum daily precipitation (in respect to the average: up to 4 mm and above 64 mm, respectively) to the yearly total is increasing (Homar et al., Reference Homar, Ramis, Romero and Alonso2010). During the last few decades, intense rain events in the western Mediterranean Basin have caused numerous catastrophic floods in several countries, with human casualties (Pastor et al., Reference Pastor, Estrela, Peñarrocha and Millán2001), like those during autumn 2018 in Mallorca Island (Grimalt-Gelabert et al., Reference Grimalt-Gelabert, Rosselló-Geli and Bauzà-Llinàs2020, Reference Grimalt-Gelabert, Bauzà-Llinás and Genovart-Rapado2021). Studying past extreme events can be useful in dealing with the coming challenges as consequences of global warming. Considering the climate sensitivity of this region, numerous studies based on speleothems have reconstructed the climate of different periods in the past (Hodge, Reference Hodge2004; Hodge et al., Reference Hodge, Richards, Smart, Ginés and Mattey2008; Dumitru et al., Reference Dumitru, Onac, Polyak, Wynn, Asmeron and Fornós2018; Torner et al., Reference Torner, Cacho, Moreno, Sierro, Martrat, Rodriguez-Lazaro and Frigola2019; Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021) as well as past sea levels (Vesica et al., Reference Vesica, Tuccimei, Turi, Fornós, Ginés and Ginés2001; Polyak et al., Reference Polyak, Onac, Fornós, Hay, Asmeron, Dorale, Ginés, Tuccimei and Ginés2018; Dumitru et al., Reference Dumitru, Austermann, Polyak, Fornós, Asmerom, Ginés, Ginés and Onac2019).

During the Little Ice Age (LIA; 1275 to 1850 CE in Cisneros et al. [2016]), frequent extreme rain events have been described in previous global studies. For instance, some of these extreme events in the western Mediterranean region were: an increase in the runoff in the Alboran Sea (Nieto-Moreno et al., Reference Nieto-Moreno, Martínez-Ruiz, Giralt, Jiménez-Espejo, Gallego-Torres, Rodrigo-Gámiz, García-Orellana, Ortega-Huertas and de Lange2011) and in the central-western part (Margaritelli et al., Reference Margaritelli, Cisneros, Cacho, Vallefuoco, Rettori and Lirer2018); humid episodes in the northern part (Bassetti et al., Reference Bassetti, Berné, Sicre, Dennielou, Alonso, Buscail, Jalali, Hebert and Christophe Menniti2016); predominant wet conditions in the Iberian Peninsula and northern Morocco coincident with frequent North Atlantic Oscillation (NAO) phases (Ait Brahim et al., Reference Ait Brahim, Wassenburg, Cruz, Sifeddine, Scholz, Bouchaou, Dassi_e, Jochum, Edwards and Cheng2018; Ramos-Román et al., Reference Ramos-Román, Jiménez-Moreno, Camuera, García-Alix, Anderson, Jiménez-Espejo and Carrión2018); the most important flood of the last millennium in the Ebro River basin (Balasch et al., Reference Balasch, Pino, Ruiz-Bellet, Tuset, Barriendos, Castelltort and Peña2019); increased lake levels in southern Spain (Martín-Puertas et al., Reference Martín-Puertas, Jiménez-Espejo, Martínez-Ruiz, Nieto-Moreno, Rodrigo, Mata and Valero-Garcés2010); and flood event enhancement in the Iberian Peninsula (Barriendos and Martin-Vide, Reference Barriendos and Martin-Vide1998; Benito et al., Reference Benito, Sopeña, Sánchez-Moya, Machado and Pérez-González2003; Moreno et al., Reference Moreno, Valero-Garcés, González-Sampériz and Rico2008; Barriendos et al., Reference Barriendos, Gil-Guirado, Pino, Tuset, Pérez-Morales, Alberola and Costa2019). The general LIA climate variability has been attributed to repeated volcanic eruptions in a short time (Crowley, Reference Crowley2000; Robock, Reference Robock2000; Bertler et al., Reference Bertler, Mayewski and Carter2011; Miller et al., Reference Miller, Geirsdóttir, Zhong, Larsen, Otto-Bliesner, Holland and Bailey2012; McGregor et al., Reference McGregor, Evans, Goosse, Leduc, Martrat, Addison and Mortyn2015), solar minima (Bard et al., Reference Bard, Raisbeck, Yiou and Jouzel2000; Mayewski et al., Reference Mayewski, Maasch, Yan, Kang, Meyerson, Sneed and Kaspari2006; Ammann et al., Reference Ammann, Joos, Schimel, Otto-Bliesner and Tomas2007), and changes in the thermohaline circulation (Broecker, Reference Broecker2000, Reference Broecker2001; Lund et al., Reference Lund, Lynch-Stieglitz and Curry2006).

Regarding the techniques used in this study, micro-computed tomography (micro-CT) and X-ray fluorescence (XRF) core scanning (henceforth XRFCS) allow rapid, continuous, nondestructive, repetitive, and high-resolution analyses of sedimentary sequences (Frigola et al., Reference Frigola, Canals and Mata2015). Consequently, these techniques are very valuable techniques in earth science research (Mees et al., Reference Mees, Swennen, Van Geet and Jacobs2003). Micro-CT is a 3D imaging and analysis method to investigate internal structures of a large variety of objects, including geomaterials (Cnudde and Boone, Reference Cnudde and Boone2013). XRFCS allows analysis of the qualitative elemental composition for elements between aluminium and uranium with resolutions up to 100 μm (Jansen et al., Reference Jansen, Van der Gaast, Koster and Vaars1998; Rothwell and Rack, Reference Rothwell, Rack and Rothwell2006). This technique is widely applied in geosciences studies to better characterize the sources of and processes responsible for the final deposits (Moreno et al., Reference Moreno, Cacho, Canals, Gromalt and Sanchez-Vidal2004; Morellón et al., Reference Morellón, Valero-Garcés, González-Sampériz, Vegas-Vilarrúbia, Rubio, Rieradevall and Delgado-Huertas2011; Nieto-Moreno et al., Reference Nieto-Moreno, Martínez-Ruiz, Giralt, Jiménez-Espejo, Gallego-Torres, Rodrigo-Gámiz, García-Orellana, Ortega-Huertas and de Lange2011; Rodrigo-Gámiz et al., Reference Rodrigo-Gámiz, Martínez-Ruiz, Jiménez-Espejo, Gallego-Torres, Nieto-Moreno, Romero and Ariztegui2011; Frigola et al., Reference Frigola, Canals and Mata2015; Cisneros et al., Reference Cisneros, Cacho, Frigola, Canals, Masqué, Martrat and Casado2016; Torner et al., Reference Torner, Cacho, Moreno, Sierro, Martrat, Rodriguez-Lazaro and Frigola2019; Cerdà-Domènech et al., Reference Cerdà-Domènech, Frigola, Sanchez-Vidal and Canals2020).

In reference to the application of both techniques to speleothems, the number of studies is scarce but growing, particularly in recent years. Since the pioneering work of Mickler et al. (Reference Mickler, Ketcham, Colbert and Banner2004), which explored the potential of micro-CT to determine the growth axis, this technique has been applied in speleothems with diverse objectives: studying 3D textures of speleothems, like those that develop inside historic walls (Martínez-Martínez et al., Reference Martínez-Martínez, Fusi, Barberini, Cañaveras and Crosta2010); exploring the potential of the stalagmite for fluid-inclusion analyses (Zisu et al., Reference Zisu, Schwarcz, Konyer, Chow and Noseworthy2012); obtaining paleoclimate proxies in unsectioned stalagmites (Walczak et al., Reference Walczak, Baldini, Baldini, McDermott, Marsden, Standish, Richards and Andreo2015); characterizing petrological features (Vanghi et al., Reference Vanghi, Iriarte and Aranburu2015); and explaining U-Th outliers (Bajo et al., Reference Bajo, Hellstrom, Frisia, Drysdale, Black, Woodhead and Borsato2016). The XRFCS technique has been applied in speleothems for reconstructing variations in atmospheric sulphate (Frisia et al., Reference Frisia, Borsato, Fairchild and Susini2005), paleoclimate and paleoenvironmental changes (Dandurand et al., Reference Dandurand, Maire, Ortega, Devès, Lans, Morel and Perroux2011; Martínez-Pillado et al., Reference Martínez-Pillado, Iriarte, Álvaro, Ortega, Aranburu and Arsuaga2020), volcanic eruptions (Badertscher et al., Reference Badertscher, Borsato, Frisia, Cheng, Edwards, Tüysüz and Fleitmann2014), and paleoflood events in caves (Finné et al., Reference Finné, Kylander, Boyd, Sundqvist and Löwemark2015; Denniston and Luetscher, Reference Denniston and Luetscher2017) and as a tool for speleothem trace element analysis (Vansteenberge et al., Reference Vansteenberge, Winter, Sinnesael, Xueqin, Verheyden and Claeys2020).

Speleothems have also been used as proxies for extreme rainfall or other hydrologic drivers of cave flooding (Gázquez et al., Reference Gázquez, Calaforra, Forti, Ghaleb and Delgado-Huertas2014; González-Lemos et al., Reference González-Lemos, Müller, Pisonero, Cheng, Edwards and Stoll2015; Denniston and Luetscher, Reference Denniston and Luetscher2017). For instance, elevated Ti data (obtained by XRFCS) compared with the calcite matrix in speleothems have been used to identify paleofloods events in southern Greece (Finné et al., Reference Finné, Kylander, Boyd, Sundqvist and Löwemark2015). In addition, previous studies have traced cave floods in the form of detrital layers recorded in stalagmites (Borsato et al., Reference Borsato, Quinif, Bini and Dublyansky2003; Dorale et al., Reference Dorale, Lepley and Edwards2005; Dasgupta et al., Reference Dasgupta, Saar, Edwards, Shen, Cheng and Alexander2010).

In the present study, the micro-CT and XRFCS techniques are coupled to explore a discontinuity in the studied speleothem (Séan stalagmite, Sa Balma des Quartó cave, Mallorca). After combining the results with geochemical data and fabric observations of the stalagmite previously published in Cisneros et al. (Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021), we present here the first description of a cave flooding event based on a speleothem from the central-western Mediterranean region.

CAVE AND CLIMATIC SETTINGS

The Seán stalagmite was recovered from Sa Balma des Quartó cave, located in the southeastern part of Mallorca in the Balearic Islands (see Fig. 1a and b).The cavity is formed in upper Miocene reef limestone (Ginés et al., Reference Ginés, Fornós, Ginés, Merino and Gràcia2014) and is located near a cliff that reaches a maximum height of 20 m, running parallel to the coast. The Miocene calcarenites are rich in marine microfossils. The length of the cave is 70 m and consists basically of a wide chamber formed by collapse (Fig. 1c and d), characterized by a rich decoration of speleothems (Bermejo et al., Reference Bermejo, Mateu, López, Minguillón, Herráez and Villar2014). The bottom of the chamber is 12.5 m from the surface and 10.5 m above sea level. The access to the cave consists of a small vertical hole in the outside rock shelter. According to the monitoring results (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021), the cave annual ventilation begins to be enhanced by the end of autumn. During winter and spring is when the cave is more ventilated. At present, the surrounding area consists of a forest formed by Mediterranean vegetation (species like Quercus ilex, Pistacia lentiscus, or genus Juniperus; Bolós, Reference Bolós1996), which is rather undisturbed by human activities.

Figure 1. Studied area. (a) Map of the western Mediterranean showing Balearic Islands, where Mallorca is located. (b) Distribution of the superficial hydrology of Mallorca Island; streams are indicated in blue and watersheds in red (adapted from Grimalt-Gelabert et al., Reference Grimalt-Gelabert, Rosselló-Geli and Bauzà-Llinàs2020). Location of Sa Balma des Quartó cave is indicated as well as the municipalities most affected by the severe flood that occurred on October 9, 2018 (Grimalt-Gelabert et al., Reference Grimalt-Gelabert, Bauzà-Llinás and Genovart-Rapado2021). (c) Topography of Sa Balma des Quartó cave (Bermejo et al., Reference Bermejo, Mateu, López, Minguillón, Herráez and Villar2014). Position of the vertical profile in part d is also indicated. Blue circles correspond to the location of the speleothems recovered in situ. (d) Vertical profile (“G-g” section in Bermejo et al., Reference Bermejo, Mateu, López, Minguillón, Herráez and Villar2014). Filled circles in c and d represent sites where evidence of cave flooding was observed after the extreme rain event of October 2018.

Sa Balma des Quartó cave is dominated by seepage flow (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021). Despite that, evidence of episodic water flows has been observed in one side of the cave after the extreme rain event that occurred during autumn 2018 in Mallorca Island (Grimalt-Gelabert et al., Reference Grimalt-Gelabert, Bauzà-Llinás and Genovart-Rapado2021, Reference Grimalt-Gelabert, Rosselló-Geli and Bauzà-Llinàs2020; Fig. 2).

Figure 2. Pictures of evidence of flood into Sa Balma des Quartó cave in November 2018: (a and b) in the upper part of the cave, gours containing water (blue lines); (c) in the lower part of the cave, marks of water flows (purple lines).

Climate in Balearic Islands corresponds to a typical Mediterranean pattern, being characterized by mild, wet winters and warm to hot, dry summers (Lionello et al., Reference Lionello, Malanotte-Rizzoli and Boscolo2006). The maximum rainfall occurs during the autumn and decreases during the winter and spring, with a very dry summer season. In Mallorca Island, present-day rainfall is characterized by high irregularity, with alternation of dry and wet years, and is affected by intense precipitation events related to cyclogenic Mediterranean conditions and the geographic trends of the island (the precipitation in the northern mountains is usually higher than in the southern zones; Grimalt-Gelabert et al., Reference Grimalt-Gelabert, Rosselló-Geli and Bauzà-Llinàs2020). Heavy rainstorms, which can occasionally produce 400 mm of rain in 24 hours, are typical and affect mainly the mountainous area in the northwestern part of the island, as well as on the east coast (Grimalt and Rosselló, Reference Grimalt, Rosselló, Antronico and Marincioni2018), where the studied cave is located (Fig. 1b).

The rainfall distribution on the western Mediterranean region has been associated with atmospheric circulation patterns as the NAO. Positive NAO modes cause stronger winter storms crossing the Atlantic on a more northerly track, reducing the transport of humidity over the Mediterranean and enhancing the evaporation–precipitation balance (Tsimplis and Josey, Reference Tsimplis and Josey2001; Lionello and Sanna, Reference Lionello and Sanna2005).

MATERIAL AND METHODS

This study aims to characterize a discontinuity in the Seán stalagmite by means of two nondestructive techniques: (1) high-resolution micro-CT and (2) XRFCS. Micro-CT was applied in a section of the Seán stalagmite (4.75–7.75 cm) where the discontinuity was observed and where a thin section had been previously performed. XRFCS was used in the same section to obtain the qualitative elemental composition and colour measurements by means of a high-resolution colour line-scan camera.

The data derived from the nondestructive techniques, which are presented for the first time in this study, are combined with geochemical data (δ18O, δ13C, trace element ratios, and U/Th dates) and fabric observations performed on the Seán stalagmite and published in Cisneros et al. (Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021). A second stalagmite from the same cave, named Multieix (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021), is taken into account at some points of the discussion, as it overlaps with the Seán stalagmite.

Seán stalagmite and previous data

The Seán stalagmite was found broken in the lower part of Sa Balma des Quartó cave (Fig. 1). Before analysis, the stalagmite was cut into two halves, and the fabric and the discontinuities along the axial length were determined using optical microscopy of thin sections (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021).

The internal microstratigraphy of the Seán stalagmite (approx. 12 × 5 cm) was characterized following the methodology and the nomenclature proposed by Martín-Chivelet et al. (Reference Martín-Chivelet, Muñoz-García, Cruz, Ortega and Turrero2017). Seán has a regular and flat external surface and a cylindrical shape and presents a white and translucent surface along the growth axis (Fig. 3). It has been divided in two unconformity-bounded units (UBUs): UBU2 from the base to 6.00 cm and UBU1 from 6.00 cm to the top. In a general way, the morphostratigraphic units could be described as flat-topped with well-defined and subhorizontal growth layers, and aggradational stacking pattern sets (Muñoz-García et al., Reference Muñoz-García, Cruz, Martín-Chivelet, Ortega, Turrero and López-Elorza2016) are generalized along this stalagmite. Most of the speleothem is characterized by columnar calcite fabrics and a remarkable and uniform lamination (<1 mm) along the growth axis. For a more detailed characterization of the entire stalagmite, refer to the supplementary material in Cisneros et al. (Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021).

Figure 3. Synthetized description of the Seán stalagmite (from Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021). The brown dashed line indicates the layer/unconformity ca. 6.00 cm. Shown in relation to the fabric pictures are those around 6.00 cm (from top to bottom): top, isolated mosaic fabrics (Mo) above the micrite layer (m), columnar fabric below (C), cross-polarized light; bottom, columnar (C), mosaic (Mo), and micrite (m) fabrics, plane-polarized light. U-Th age model performed on Bchron. Red plots are the final age models. Black diamonds represent the U-Th ages (2σ error). Grey dashed lines correspond to the total range of ages covered by all the age models obtained with Bchron, which have statistically the same significance (95&per; confidence interval).

The age of the Seán stalagmite is based on U-Th dating of nine samples, published in Cisneros et al. (Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021; Table 1) and performed according to the procedures of Edwards et al. (Reference Edwards, Chen and Wasserburg1987). Ages were obtained with a Neptune Thermo Finnigan Multi-Collector ICP-MS at the University of Minnesota (USA). Ages are corrected taking into account an initial 230Th/232Th atomic ratio of (4.4 ± 2.2) × 10−6 (Cheng et al., Reference Cheng, Edwards, Chou Shen, Polyak, Asmerom, Woodhead and Hellstrom2013).

Table 1. Summary of the results of U-Th analyses used in the age model of the Seán stalagmite (2σ error) and published in Cisneros et al. (Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021).

a Age in italics is not included in the age model. All years are CE.

The age model was produced using the R statistics package Bchron (Parnell et al., Reference Parnell, Haslett, Allen, Buck and Huntley2008). It was performed in two sections to minimize uncertainties: from the bottom to the unconformity at 6.00 cm and from there to the top. One U-Th age was excluded to minimize the outlier probability in the adjacent dates (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021).

The lower and upper boundaries of the age uncertainties, according to all the age models obtained by Bchron, have statistically the same significance (95%). The results show that Seán grew quite continuously between 1421 CE and 1880 CE with a mean growth rate of 0.4 mm/yr (from 0.1 to 0.9 mm/yr).

UBU1 and UBU2 are separated by a brown millimetric layer ca. 6.00 cm that seems to correspond to a detrital or allogenic horizon. At 6.00 cm from the top, this millimetric brown layer shows micrite fabric, and above this layer, some isolated mosaic fabrics have been also observed. Around 6.00 cm, skeletal components have been also distinguished, which could correspond to a gastropod of 1400 μm length. (Fig. 4). Although the brown layer follows the general lamination pattern, its thickness varies between 1.3 and 2.5 mm. The thickness of the detrital horizon is 0.25 mm (located between 6.25 and 6.00 cm). According to the age model (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021), the distinctive horizon was accumulated after 1616 ± 23 CE (6.25 cm) and before 1623 ± 28 CE (6.00 cm). Because the growth rates around the micrite layer were similar to the average growth rates obtained from the UBU2 part, the previous study concluded that no significant erosion took place (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021; Fig. 3).

Figure 4. Fabric picture of the Seán stalagmite thin section ca. 6.00 cm. Columnar (C), mosaic (Mo), and micrite (m) fabrics, plane-polarized light. The arrow indicates the gastropod test.

Stable isotopes and trace element samples were manually microdrilled at 1 mm intervals along the growth axis. Although δ18O data in the centimetres around the brown layer are similar to the mean values of the entire stalagmite record (from 4.7 to −4.9 ‰), δ18O values right along the detrital horizon tend to be more depleted (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021). δ13C values just before the beginning of the layer are enriched considering the rest of the record (6.25 cm, −6.50 ‰), and values tend to be more depleted along the layer (−8.2 ‰). In reference to trace elements (Mg/Ca, Sr/Ca, and Ba/Ca), the values of the ratios present fluctuations during the detrital horizon, and the general trend is to decrease (from 7.33 to 5.47, from 0.13 to 0.083, from 0.009 to 0.006, respectively.

To gain a better understanding of the event that affected the cave and caused the brown layer in Seán, a second stalagmite from the same cave has been included in the “Discussion.” This stalagmite, named Multieix, was found ~5 m distant from where the Seán stalagmite was lying (Fig. 1c). The particular morphology and length of this stalagmite, which presents three different growth axes, has been also observed in remaining active stalagmites in a specific site within the cave, just where it was found. This suggests that it was recovered near its growth position, although its base was not identified. Considering all of this, Multieix's growth position is located approximately 2 m higher than the Seán site. The Multieix stalagmite is 30.5 cm in length and grew from the year 302 BCE to 1844 CE. Only its top 5 cm grew up simultaneously with Seán (during the 1503–1844 CE period). No new data for the Multieix stalagmite are presented in this study; more detailed information can be found in Cisneros et al. (Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021).

Micro-CT scanning

The density of the stalagmite has been studied in Seán ca. 6.00 cm (~4.75–7.75 cm) using micro-CT scanning with a Multitom CORE X-ray CT system in the CORELAB, University of Barcelona. The scanner was operated at a tube voltage of 140 kV and an intensity of 33 W. The exposure time was 800 ms, the voxel size was 33 μm, and the number of projections was 1500. Reconstructed images of the relative density values were exported as 16-bit images using the AVIZO (FEI) software. This technique allows obtaining slices with voxel resolution ranging from 5 to 300 μm (Frigola et al., Reference Frigola, Canals and Mata2015). The micro-CT scanning results are expressed in this study as 2D images in a colour map scale denoting relative density values.

XRFCS

The qualitative elemental composition of the Seán stalagmite was studied between ~4.75 and 7.75 cm from the top, corresponding to the same segment used for micro-CT analyses, with an Avaatech XRF core-scanner system (CORELAB, University of Barcelona). The scanning resolution was 200 μm, and the measured window was adjusted to 8 mm to align with the laminae of the stalagmite. Analyses were focused on the left part of the growing axis. Elements with atomic weights between aluminium and iron were analysed at 10 kV, 1.95 mA, and a 60 s exposure time, while elements with atomic weights between iron and lead were analysed at 30 kV, 1.95 mA, and a 90 s exposure time.

The colour measurements were conducted along the left side of the growth axis of the entire Seán stalagmite. A high-resolution (70 micron pixels) visible image of the Seán stalagmite was obtained with a colour line-scan camera mounted in the Avaatech XRFCS System. The acquisition software provides images in several formats and also numerical colour values in RGB and CIE-Lab coordinates at the same high resolution. Visible pictures do not require any additional treatment, as is shown in Figure 3. The L*a*b* values (colour coordinates) are at the scale of uniform colour defined by the Commission Internationale de l'Eclairage (CIE; Muñoz et al., Reference Muñoz, Bartolomé, Muñoz, Sancho, Moreno, Hellstrom, Osácar and Cacho2015). L* corresponds to lightness and ranged in value from 0 (black) to 100 (white). The proxies a* and b* represent variations between red-green and yellow-blue, respectively, ranging in values between −120 and 120 (Westland, Reference Westland2012). Colour coordinates enable recognition of changes in the carbonate content, as well as the presence of iron and clays (Mix et al., Reference Mix, Rugh, Pisias and Veirs1992; Nederbragt and Thurow, Reference Nederbragt, Thurow and Francus2004; Rogerson et al., Reference Rogerson, Weaver, Rohling, Lourens, Murray, Hayes and Rothwell2006; Debret et al., Reference Debret, Sebag, Desmet, Balsam, Copard, Mourier and Susperrigui2011).

RESULTS

Micro-CT scanning

Micro-CT scanning provides a density colour map that reveals certain differences in the analysed section at around 6.00 cm of the Seán stalagmite (Fig. 5, left panel). Along the growth axis, a continuous lamination has been observed, consisting of dense laminae (25,081–15,673.5 Hounsfield units [HU]) alternating with more porous laminae (15,673.5–6266 HU).

Figure 5. Micro-computed tomography (micro-CT) and X-ray fluorescence (XRF) core scanning (XRFCS) results (this study) from the boundary of the two unconformity-bounded units (UBUs) in Seán (4.75–7.75 cm). Left, Colour map from micro-CT between (HU, Hounsfield units); horizontal grey bar indicates the detrital layer (from 6.00 cm to 6.25 cm). Right panel (from left to right), Coordinates of colour CIE L (Lightness)*a*b*; S, Ti, Ca, and Mn elements (expressed as peak area). XRFCS measurements were carried out on the left part of the growth axis of the stalagmite. Vertical and discontinuous black line (above the left image) indicates the line of measurements.

Colour-map densities < 6266 HU correspond to air measurements, and the 15,673.5 HU value to dispersions of the X-ray beam due to the edge effect. The brown lamina at ca. 6.00 cm presents high air content (~6266 HU), and below this lamina, certain lineal structures are observed, which are not present in UBU1. The 2D images point out a high air content in the brown layer across the entire stalagmite and not only in the surface of the split stalagmite. Thus, in agreement with Vanghi et al. (Reference Vanghi, Iriarte and Aranburu2015), micro-CT has improved the characterization of the speleothem, providing information about spatial distribution of porosity and stratigraphic architecture.

XRFCS results

Colour measurements and qualitative elemental composition results obtained for the boundary between the two UBUs of Seán at ca. 6.00 cm are shown in Figure 5. Colour measurements (L*a*b* values for colour coordinates) obtained from the high-resolution pictures define an abrupt change in the boundary between both UBUs of Seán (ca. 6.00 cm). Parameters b* (yellow-blue) and L* (lightness) indicate more differences in the boundary than parameter a* (red-green). The three parameters increase their values along the brown layer starting at 6.25 cm until 5.90 cm.

Regarding XRFCS results, where S, Ti, Mn, and Ca profiles are expressed as peak areas (Fig. 5), the examined section displays significant oscillations in the analysed elements. All of these elements exhibit a certain enhancement beginning around 6.50 cm. S values decrease around 6.75 cm and increase around 6.00 cm. The S record shows several oscillations but without any significant change associated with the UBU limit. Ti shows an abrupt enhancement from 6.20 cm until 6.00 cm (~1617 ± 25 CE to 1623 ± 28 CE), reaching the maximum values at 6.00 cm. In the Mn profile, a remarkably rapid increase occurs at ca. 4.70 cm (1671 ± 13 CE), and the lowest values were obtained at 5.60 cm (1642 ± 26 CE). This element also shows a relative enhancement at ca. 6.00 cm. With regard to Ca, maxima values occurred at 6.80 cm (1585 ± 16 CE), and around 6.00 cm, this element shows a relatively less abrupt increase than the other elements presented here. At the beginning of the brown layer (6.25 cm), Ca values are low (~480,000 peak area) and tend to decrease and, later, to be enhanced until 6.00 cm (~554,000 peak area).

DISCUSSION

Interpretation of the allogenic or detrital horizon: sediment sources and causes

The micrite layer at ca. 6.00 cm marks the limit between the two described UBUs of the Seán stalagmite (Fig. 5). This layer is associated with a significant Ti peak and is also accompanied by abrupt changes in the colour parameters, according to the results obtained by the XRFCS (Fig. 5). The brown layer follows the general lamination according to micro-CT scanning results; no important signs of erosion or hiatuses in the deposition have been detected; and thus, Seán apparently grew quite continuously. Although a microhiatus cannot be absolutely disregarded, this unconformity consists of a minor event that does not seem to be responding to longer-term variations in drip rates or drastic changes in the growth rates.

In the study by Finné et al. (Reference Finné, Kylander, Boyd, Sundqvist and Löwemark2015) in southern Greece, elevated Ti (obtained by XRFCS) compared with the calcite matrix has been used to identify paleoflood events. In the Seán stalagmite, the presence of high Ti values and also of a micrite layer suggests the enhanced arrival of terrigenous particles, corresponding to an allogenic or detrital horizon.

In this study, among the potential sources of this distinctive layer, the influence of a large volcanic eruption was initially considered, because such events have frequently been attributed to the dominant cold temperatures of the LIA (Crowley, Reference Crowley2000; Robock, Reference Robock2000; Bertler et al., Reference Bertler, Mayewski and Carter2011; Miller et al., Reference Miller, Geirsdóttir, Zhong, Larsen, Otto-Bliesner, Holland and Bailey2012; McGregor et al., Reference McGregor, Evans, Goosse, Leduc, Martrat, Addison and Mortyn2015). Nevertheless, this layer is not related to any sulphur (S) peak or δ13C enrichment in the Seán stalagmite (Fig. 6d and f), both of which are typically detected in records associated with volcanic eruptions at other locations (Frisia et al., Reference Frisia, Badertscher, Borsato, Susini, Göktürk, Cheng, Edwards, Kramers, Tüysüz and Fleitmann2008; Badertscher et al., Reference Badertscher, Borsato, Frisia, Cheng, Edwards, Tüysüz and Fleitmann2014). In addition, according to the global and Northern Hemisphere volcanic eruptions recorded by Gao et al. (Reference Gao, Robock and Ammann2008) and Crowley and Unterman (Reference Crowley and Unterman2012) for the last centuries, the period coincident with this distinctive layer of the Seán stalagmite is not characterized by the most severe eruptions recorded (Fig. 6l).

Figure 6. Results presented in this study from the Seán stalagmite compared with previously published data for the Seán and Multieix stalagmites, North Atlantic Oscillation (NAO) reconstruction and volcanism activity. (a–e) Ba/Ca, Sr/Ca, Mg/Ca, δ13C, and δ18O records for the Seán stalagmite. δ13C, δ18O records from Multieix are also shown in lighter colours (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021). (f and g) S and Ti from X-ray fluorescence (XRF) core scanning (XRFCS) analyses (this study). (h and i) Growth rates and U/Th ages (diamonds) of Seán and Multieix stalagmites (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021). (j) Mg/Ca sea-surface temperature (SST) from north Minorca (Cisneros et al., Reference Cisneros, Cacho, Frigola, Canals, Masqué, Martrat and Casado2016). (k) NAO reconstruction (Faust et al., Reference Faust, Fabian, Milzer, Giraudeau and Knies2016). (l) Northern and global volcanism (Gao et al., Reference Gao, Robock and Ammann2008; Crowley and Unterman, Reference Crowley and Unterman2012). Both subperiods of Little Ice Age (LIAa and LIAb) and the Industrial Era (IE) are also indicated. Brown vertical band indicates discontinuity and brown layer observed in Seán (~6.00 cm), which corresponds to the limit between both unconformity-bounded units (UBUs) in the Seán stalagmite.

One mechanism capable of transporting the material of the distinctive layer to our cave, which is interpreted according to the significant Ti peak obtained by means of the XRFCS in the brown layer (Fig. 6g), is the supply of dust aerosols. Numerous studies have demonstrated the influence of Saharan dust in our study area (Goudie and Middleton, Reference Goudie and Middleton2001; Moreno et al., Reference Moreno, Cacho, Canals, Prins, Sánchez-Goñi, Grimalt and Weltje2002) and its contribution to the red Mediterranean soils (Muhs et al., Reference Muhs, Budahn, Avila, Skipp, Freeman and Patterson2010). In addition, Saharan dust has been detected in cave soil sediments in Mallorca (Fornós et al., Reference Fornós, Ginés and Gràcia2009) and in stalagmites from the eastern Mediterranean through Sr and U isotopes (Frumkin and Stein, Reference Frumkin and Stein2004). Saharan dust is mostly supplied by wet deposition (“muddy rain”), when the air mass becomes fresh after being in contact with the sea, while dry deposition only takes place occasionally (Fiol et al., Reference Fiol, Fornós, Gelabert and Guijarro2005).

The period during which the distinctive horizon was accumulated according to the age model (1616 ± 23 CE to 1623 ± 28 CE) is coincident with a drop in sea-surface temperature (SST) considering the available multidecadal timescale reconstruction derived from marine sediments recovered in the study area (Cisneros et al., Reference Cisneros, Cacho, Frigola, Canals, Masqué, Martrat and Casado2016; Fig. 6j). The years of the distinctive horizon have been characterized by positive and negative NAO phases (Fig. 6k), according to the multiannual reconstruction of Faust et al. (Reference Faust, Fabian, Milzer, Giraudeau and Knies2016). If the δ18O record is interpreted as hydroclimate variability (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021), the values obtained for the period corresponding to the brown layer, which are similar to the mean values of the entire stalagmite record, do not indicate years that were especially wet or dry or that experienced extreme rain or drought events (Fig. 6e). Hence, it is not feasible to determine whether this episode corresponds to a deposition of dry or wet dust.

In agreement with previous studies, the distinctive layer and the Ti peak detected in the Seán stalagmite could be interpreted as a cave flood tracer (Borsato et al., Reference Borsato, Quinif, Bini and Dublyansky2003; Dorale et al., Reference Dorale, Lepley and Edwards2005; Dasgupta et al., Reference Dasgupta, Saar, Edwards, Shen, Cheng and Alexander2010; Finné et al., Reference Finné, Kylander, Boyd, Sundqvist and Löwemark2015) and/or as a proxy for extreme rainfall or other hydrologic drivers of cave flooding (Gázquez et al., Reference Gázquez, Calaforra, Forti, Ghaleb and Delgado-Huertas2014; Denniston and Luetscher, Reference Denniston and Luetscher2017).

Although Sa Balma des Quartó cave is dominated by seepage flow (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021) and has been described as a very dry cave in the present day, evidence of episodic water flows has been observed in one side of the cave during October 2018 after an extreme rain event that may have led to the inundation of the base of the cave (Figs. 1d and 2). An inundation like that one occurring in the past could easily have covered the 6.00 cm of Seán studied here, which would be the length of the stalagmite at the moment of the flooding (Fig. 7). The hypothesis of the cave flooding agrees with the fact that the micrite layer fills the gaps between the older columnar fabric (Fig. 4).

Figure 7. Hypothetical reconstruction of the cave flooding during the seventeenth century. Topography corresponds to the vertical profile of Sa Balma des Quartó cave (“G-g” section in Bermejo et al., Reference Bermejo, Mateu, López, Minguillón, Herráez and Villar2014) also shown in Fig. 1d. Blue circles correspond to the locations where the speleothems were found, which are separated by a distance of ~5 m. Seán's growth position is located approximately 2 m lower than the Multieix site. The hypothetical flood level, which is represented in qualitative terms, would have covered the 6.00 cm length of the Seán stalagmite in that moment. The evidence of cave flooding observed after the extreme rain event of October 2018 (gours with water in the upper part and water flow marks in the lower part) is also indicated.

When considering the distinctive layer of the Seán stalagmite as an allogenic or detrital horizon, distinction between an absolute external source of the sediment (allogenic material) or absolute in situ fine sediments mobilized by water (detrital material) cannot be established. However, the presence of the gastropod in the distinctive layer of the stalagmite (Fig. 4) likely points to the mobilisation by water of the in situ sediments. The Miocene calcarenites of the karst system are rich in these marine microfossils, which could arrive in suspension to the stalagmite and be fixed by the next calcite growth. The arrival of the gastropod to the surface of the stalagmite via drip water or by falling down from the limestone of the cave ceiling cannot be excluded.

Present and past cave floodings and extreme rain events

On October 9, 2018, an extreme rain event occurred in Mallorca, resulting in a severe flood due to the high intensity and amount of precipitation. The flood took place on the eastern coastline of the island, affecting the municipalities of Sant Llorenç des Cardassar, Artà, and to a lesser extent, Capdepera and Manacor, and was the most severe flood occurring in Sant Llorenç in the last 80 yr (Grimalt-Gelabert et al., Reference Grimalt-Gelabert, Bauzà-Llinás and Genovart-Rapado2021; Fig. 1b). The amount of rainfall in Sant Llorenç during October 9, 2018, was 257 mm, and the effects were severe: nine lives were lost, several bridges and infrastructures collapsed, and numerous vehicles were dragged (Grimalt-Gelabert et al., Reference Grimalt-Gelabert, Rosselló-Geli and Bauzà-Llinàs2020, 2021). Several weeks later, on 16 November, we visited Sa Balma des Quartó cave and found clear evidence of cave flooding as a consequence of the extreme event. The group of gours in the upper part of the cave were full of water for the first time in the 5 yr of monitoring (Fig. 2), and water flow tracks were observed in the cave sediments of the lower part of the cave (Figs. 1d and 2). Typically, Sa Balma des Quartó cave has been described as a very dry cave in the present day, and the gours of the upper part of the cave have been described as being empty of water (Bermejo et al., Reference Bermejo, Mateu, López, Minguillón, Herráez and Villar2014). Regarding the floor of the lower part of the cave, it is covered by sediments and acts like a sediment sink (Bermejo et al., Reference Bermejo, Mateu, López, Minguillón, Herráez and Villar2014); observation of flow tracks in the sediments have not been frequent during the last years.

These observations inside the cave indicate a response associated with extreme rainfall events and reinforce the hypothesis that the micrite layer of the Seán stalagmite can be interpreted as a result of past cave flooding, which occurred after 1616 ± 23 CE (6.25 cm) and before 1623 ± 28 CE (6.00 cm). The cave was possibly flooded by an extreme rain event, with enough water to cover the 6.00 cm length of the Seán stalagmite. The water could flow over the detrital sediments deposited earlier in the cave, supplying these sediments to the stalagmite surface (Fig. 7). Nevertheless, it cannot be discarded that the water inflow could introduce sediments from the external surface into the cave. It should be noted that the length of the Seán stalagmite regarding the current position of the cave floor could possibly be different from today due to the sediments supplied by the inflow water.

The calcite growth previous to the micrite layer shows columnar fabrics and micro-CT scanning results that point out significant high relative density values (>15,673.5 HU), as was expected to find associated with this kind of fabric (Vanghi et al., Reference Vanghi, Iriarte and Aranburu2015; Fig. 5). Columnar fabrics have usually been related to constant drip rates (Frisia, Reference Frisia2015) that in some way could be associated with rather constant climatic conditions. These conditions could be interpreted to be predominant during a few years before 1615 ± 23 CE.

The micrite layer presents a high air content according to the micro-CT data and a significant abrupt Ti increase, as well as a slight enhancement of Mn and Ca, according to the XRFCS data. In this micrite layer, δ18O values tend to be more depleted and the values of the trace element ratios tend to decrease (Fig. 6a–e). The more depleted δ18O values could indicate a reduction of the evaporation due the contact of the stalagmite surface with the water flow as a consequence of the cave flood. Regarding the decreasing trend of Mg/Ca ratio (Fig. 6c), the relative enhancement of Ca detected by XRFCS analyses along the detrital horizon should be noted (until 6.00 cm; 1623 ± 28 CE; Fig. 5). Despite this relative Ca enhancement along the detrital horizon, Ca values tend to decrease at the beginning of the brown layer. Previous studies have recorded reduced Ca peak areas in clay horizons (Finné et al., Reference Finné, Kylander, Boyd, Sundqvist and Löwemark2015).

Concerning the contact time between the “muddy” water and the stalagmite surface, we can only assert that the posterior calcite growth, according to the age model, had already precipitated ca. 1623 ± 28 CE (~6 yr later). It should be noted that: (1) the two U/Th age uncertainties obtained close to the detrital horizon at ca. 6.00 cm (5.4 cm, 1660 ± 15 CE; and 6.6.00 cm, 1602 ± 8 CE) are really low, and (2) growth rates around the micrite layer are similar to the mean obtained in the UBU2 of the Seán stalagmite (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021; Fig. 6h and i). The time during which no layers of clean calcite (without detrital material) precipitated could indicate that the water inflow volume supplied by the rainfall was severe (Fig. 7). It should be noted that after the extreme rain event of October 2018, only a few centimetres of water were found in the gours of the upper part of the cave weeks later (Fig. 2a and b). Thus, the extreme rain that occurred in 2018 was possibly less severe than the past event of the seventeenth century, as it covered the 6.00 cm length of the Seán stalagmite (Fig. 7).

Mosaic fabrics have been observed after the micrite layer of the Seán stalagmite (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021). This kind of fabric has been reported as a product of dissolution of columnar calcite and re-precipitation of mosaic calcite driven by an influx of undersaturated waters causing dissolution of a preexisting fabric (Frisia, Reference Frisia2015). The presence of isolated crystals in mosaic after the micrite layer could possibly be associated with new contacts (occasional and/or intermittent) between the muddy water and the stalagmite surface. We consider the occurrence of the contacts as occasional and/or intermittent, because no important hiatus was detected in the record. Moreover, based on the minor event indicated by micro-CT images and fabric observations, the unconformity of the Seán stalagmite does not appear to align with significant shifts in growth rates or prolonged fluctuations in drip rates.

After the cave flooding of 2018 and those that occurred during the seventeenth century, it is likely that the stagnant water gradually dissipated through infiltration into deeper levels within the karst. It should be noted that the coeval stalagmite, Multieix, does not present any evidence of allogenic or detrital sediments around these years. This stalagmite should have a length in that moment of ~25 cm and presents a very low growth rate (Fig. 6h). Multieix's growth position is located approximately 5 m away from the Seán site but also ~2 m higher (Figs. 1c and 7). According to that, the probability that the water covered this stalagmite is much less than in the case of Seán. After the extreme rain event of 2018, no accumulated water or gours containing water were observed in the Multieix site.

The east of Mallorca, where Sa Balma de Quartó cave is located, shows the ideal conditions for episodes of heavy rain, in particular in situations of eastern cyclonic advection or Mediterranean depressions, as on the eastern coastline of the Iberian Peninsula (Grimalt-Gelabert et al., Reference Grimalt-Gelabert, Bauzà-Llinás and Genovart-Rapado2021). Previous studies based on historical documents have pointed out floods around Mallorca Island during the seventeenth century, particularly during the autumn of the years 1618, 1620, 1635, 1655, and 1683 (Campaner y Fuertes, Reference Campaner y Fuertes1881).

It should be noted that the most extreme rain event documented in the last 1000 yr occurred during November 1617 (Pino et al., Reference Pino, Alberola, Balasch, Barriendos, Gil, Grau-Satorras, Mazón, Pérez Morales and Tuset2018), when catastrophic floods have been described in more than 124 municipalities from the Mediterranean coast in the Iberian Peninsula, between Alacant and Perpignan. These floods caused general overflows and severe damage throughout the region, including the complete destruction of more than 445 infrastructures (Thorndycraft et al., Reference Thorndycraft, Barriendos, Benito, Rico and Casas2006; Pino et al., Reference Pino, Alberola, Balasch, Barriendos, Gil, Grau-Satorras, Mazón, Pérez Morales and Tuset2018). In addition, the overflows of the Llobregat and Ter Rivers in Catalonia exceeded the magnitude of the events recorded for the last 3000 yr (Thorndycraft et al., Reference Thorndycraft, Barriendos, Benito, Rico and Casas2006).

The fact that no more distinctive horizons have been detected during this period suggests the possibility of a very severe event recorded within the Seán stalagmite. While it is feasible to conduct more comprehensive research into the timing of the event, the intensity of the recorded event after 1616 ± 23 CE and before 1623 ± 28 CE likely aligns with the extreme rain event that took place in November 1617.

The micro-CT and XRFCS methodologies applied in this study have allowed a better characterization of the flood layer and its surroundings in the stalagmite and a better comprehension of the past flood event. The high relative density values (>15,673.5 HU) obtained by the micro-CT scanning in the calcite growth previous to the micrite layer (columnar fabrics) suggest that rather constant climatic conditions were predominant during a few years before 1615 ± 23 CE. Micro-CT results also indicated no important signs of erosion or hiatuses in the deposition of the Seán stalagmite, which apparently indicates continuous growth. In addition, the information provided by the 2D images and the high air content (~6266 HU) obtained in the distinctive horizon pointed out that the event, which caused the lamina, affected entire width of the stalagmite and not only the surface of the split stalagmite. According to the XRFCS data, there was an abrupt enhancement of Ti between ~1617 ± 25 CE and 1623 ± 28 CE (from 6.20 cm until 6.00 cm), confirming the enhanced arrival of terrigenous particles corresponding to the distinctive horizon.

CONCLUSIONS AND FUTURE PERSPECTIVES

In this study, we present new data of the characterization of a discontinuity in the Seán stalagmite using two nondestructive techniques: (1) high-resolution micro-CT and (2) XRFCS. Both techniques were applied to the 4.75–7.75 cm segment of the stalagmite. Micro-CT was used to study stalagmite density, while XRFCS was applied to obtain qualitative elemental composition and colour measurements by means of high-resolution colour line-scan camera. The results from these nondestructive techniques are combined with previous geochemical data (δ18O, δ13C, trace element ratios, and U/Th dates) and fabric observations published in Cisneros et al. (Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021), which were performed in the same stalagmite from Sa Balma des Quartó cave in Mallorca.

The discontinuity in the Seán stalagmite was characterized in a previous study by Cisneros et al. (Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021) as a millimetric brown layer with micrite fabric. Other observations included isolated mosaic fabrics above this layer, the presence of a gastropod, the layer filling gaps between older columnar fabric, and no significant difference in growth rates. The distinctive horizon was accumulated after 1616 ± 23 CE (6.25 cm) and before 1623 ± 28 CE (6.00 cm) and could correspond to an allogenic horizon.

The two methodologies applied in the present study have improved the characterization of the distinctive horizon. Mainly, they have provided information about spatial distribution of porosity and stratigraphic architecture (micro-CT) and qualitative composition of the layer (XRFCS), allowing a better understanding of its sediment sources and causes.

On the one hand, by means of the micro-CT scanning, the high relative density values (>15,673.5 HU) obtained below the micrite layer (columnar fabrics) suggest that rather constant climatic conditions were predominant during a few years before 1615 ± 23 CE. This technique also indicated no important signs of erosion or hiatuses in the deposition of the Seán stalagmite, which indicates that it grew continuously. In addition, the information provided by the 2D images and the high air content (~6266 HU) obtained in the distinctive horizon indicated that the event that caused the lamina affected the entire width of the stalagmite and not only the surface of the split stalagmite.. This information, together with the fact that micrite is filling the older columnar fabrics, suggests a cave flood event as a cause of the distinctive horizon. On the other hand, according to the XRFCS data, an abrupt enhancement of Ti occurred between ~1617 ± 25 CE and 1623 ± 28 CE (from 6.20 cm until 6.00 cm), which confirms the enhanced arrival of terrigenous particles, corresponding to the distinctive horizon. The terrigenous particles could correspond to an absolute external source of the sediment (allogenic material) or absolute in situ fine sediments mobilized by water (detrital material).

Although cave floodings in other regions have been identified by previous studies using detrital layers, this is the first time that such an event has been documented by an individual flood layer in the studied area and period, representing an opportunity to use speleothems to describe past extreme events in the Mallorcan climate. The hypothesis of the cave flooding during the seventeenth century is reinforced by the observations in the cave response to the 2018 extreme rainfall that occurred in Mallorca.

A second stalagmite from the same cave, named Multieix, which overlaps with the Seán stalagmite (Cisneros et al., Reference Cisneros, Cacho, Moreno, Stoll, Torner, Català, Edwards, Cheng and Fornós2021), has been used to better understand the cave flooding in the past. The absence of any allogenic/detrital horizon in this stalagmite during the seventeenth century suggests that no sediment transported by water reached its location (which was 2 m higher than Seán and had a length of 25 cm at that time). Therefore, the flood did not extend to this elevated section of the cave.

Studies based on historical documents have pointed out floods around Mallorca Island during the seventeenth century, particularly during the autumn of the years 1618, 1620, 1635, 1655, and 1683 (Campaner y Fuertes, Reference Campaner y Fuertes1881). However, the most extreme rain event documented in the last 1000 yr occurred during November 1617 (Pino et al., Reference Pino, Alberola, Balasch, Barriendos, Gil, Grau-Satorras, Mazón, Pérez Morales and Tuset2018), when catastrophic floods have been described in more than 124 municipalities from the Mediterranean coast in the Iberian Peninsula, between Alacant and Perpignan.

The fact that no more distinctive horizons have been detected in the Seán stalagmite during this period probably suggests that the event recorded after 1616 ± 23 CE and before 1623 ± 28 CE could have been very severe. Although more detailed investigation about the moment of the flood can be developed, the severity of the event recorded by the Seán stalagmite possibly corresponds to the extreme rain event of November 1617.

Thus, the two nondestructive techniques (micro-CT and XRFCS) used in this study to characterize the discontinuity of the Seán stalagmite have allowed a better characterization of the detrital layer, which strongly suggests its causal relation with a flood. Micro-CT and XRFCS techniques hold significant potential for advancing our comprehension of the mechanisms responsible for discontinuities within a stalagmite, while preserving the integrity of the stalagmite itself. These techniques can complement the information derived from other data sources, such as fabric observations, thereby enhancing our understanding of the causes and sources of the sediment forming the distinct horizons within a stalagmite.

Acknowledgments

This work was funded by the projects TIMED (683237) of the European Research Council (Consolidator Grants); LACEN-CLI (2476-S/2017); OPERA (CTM2013-48639-C2-1-R); PLIOKAR (CGL2013-48441-P); CHIMERA (CTM2016-75411-R); PLIOKAR-II (CGL2016-79246-P, AEI-FEDER, EU); PLIOKAR-III (PID2020-112720GB-I00/AEI); SPYRIT (CGL2016-77479-R); and Generalitat de Catalunya, Grups de Recerca Consolidats (2017 SGR 315) to GRC Geociències Marines. We are grateful to M. Guart and T. Bullich for their collaboration in the laboratory tasks and to J. Ginés, À. Ginés, and A. Pilares for their collaboration in the fieldwork. MC benefits from a Margarita Salas postdoctoral fellowship at the University of Barcelona from the Ministerio de Universidades-Gobierno de España funded by European Union (NextGenerationEU funds). Many thanks are given to the editors, L. Piccini, and the anonymous reviewer for their constructive comments.

References

REFERENCES

Ait Brahim, Y., Wassenburg, J.A., Cruz, F.W., Sifeddine, A., Scholz, D., Bouchaou, L., Dassi_e, E.P., Jochum, K.P., Edwards, R.L., Cheng, H., 2018. Multi-decadal to centennial hydroclimate variability and linkage to solar forcing in the Western Mediterranean during the last 1000 years. Scientific Reports 8, 17446.Google Scholar
Ammann, C.M., Joos, F., Schimel, D.S., Otto-Bliesner, B.L., Tomas, R.A., 2007. Solar influence on climate during the past millennium: results from transient simulations with the NCAR climate system model. Proceedings of the National Academy of Sciences USA 104, 37133718.Google Scholar
Badertscher, S., Borsato, A., Frisia, S., Cheng, H., Edwards, R.L., Tüysüz, O., Fleitmann, D., 2014. Speleothems as sensitive recorders of volcanic eruptions—the Bronze Age Minoan eruption recorded in a stalagmite from Turkey. Earth and Planetary Science Letters 392, 5866.Google Scholar
Bajo, P., Hellstrom, J., Frisia, S., Drysdale, R., Black, J., Woodhead, J., Borsato, A., et al., 2016. “Cryptic” diagenesis and its implications for speleothem geochronologies. Quaternary Science Reviews 148, 1728.Google Scholar
Balasch, J.C., Pino, D., Ruiz-Bellet, J.L., Tuset, J., Barriendos, M., Castelltort, X., Peña, J.C., 2019. The extreme floods in the Ebro River basin since 1600 CE. Science of the Total Environment 646, 645660.Google Scholar
Bard, E., Raisbeck, G., Yiou, F., Jouzel, J., 2000. Solar irradiance during the last 1200 years based on cosmogenic nuclides. Tellus B 52, 985992.Google Scholar
Barriendos, M., Gil-Guirado, S., Pino, D., Tuset, J., Pérez-Morales, A., Alberola, A., Costa, J., et al., 2019. Climatic and social factors behind the Spanish Mediterranean flood event chronologies from documentary sources (14th–20th centuries). Global and Planetary Change 182, 102997.Google Scholar
Barriendos, M., Martin-Vide, J., 1998. Secular climatic oscillations as indicated by catastrophic floods in the Spanish Mediterranean coastal area (14th–19th centuries). Climatic Change 38, 473491.Google Scholar
Bassetti, M.A, Berné, S., Sicre, M.A., Dennielou, B., Alonso, Y., Buscail, R., Jalali, J., Hebert, B., Christophe Menniti, C., 2016. Holocene hydrological changes of the Rhone River (NW Mediterranean) as recorded in the marine mud belt. Climate of the Past 12, 15391553.Google Scholar
Benito, G., Sopeña, A., Sánchez-Moya, Y., Machado, M.J., Pérez-González, A., 2003. Palaeoflood record of the Tagus River (central Spain) during the Late Pleistocene and Holocene. Quaternary Science Reviews 22, 17371756.Google Scholar
Bermejo, J., Mateu, T., López, B., Minguillón, R., Herráez, G., Villar, A., 2014. Cova de sa Balma des Quartó (Manacor, Mallorca). Endins: publicació d'espeleologia, no. 36, 5964.Google Scholar
Bertler, N.A.N, Mayewski, P.A., Carter, L., 2011. Cold conditions in Antarctica during the Little Ice Age: implications for abrupt climate change mechanisms. Earth and Planetary Science Letters 308, 4151.Google Scholar
Bolós, O. de, 1996. La vegetació de les Illes Balears. Institut d'Estudis Catalans, Barcelona, p. 269.Google Scholar
Borsato, A., Quinif, Y., Bini, A., Dublyansky, Y., 2003. Open-system alpine speleothems: implications for U-series dating and paleoclimate reconstructions. Studi Trentini di Scienze Naturali, Acta Geologica 80, 7183.Google Scholar
Broecker, W.S., 2000. Was a change in thermohaline circulation responsible for the Little Ice Age? Proceedings of the National Academy of Sciences USA 97, 13391342.Google Scholar
Broecker, W.S., 2001. Paleoclimate: was the medieval warm period global? Science 291, 14971499.Google Scholar
Campaner y Fuertes, A., 1881. Cronicón Mayoricens. J. Colomer y Salas, Palma.Google Scholar
Cerdà-Domènech, M., Frigola, J., Sanchez-Vidal, A., Canals, M., 2020. Calibrating high resolution XRF core scanner data to obtain absolute metal concentrations in highly polluted marine deposits after two case studies off Portmán Bay and Barcelona, Spain. Science of the Total Environment 717, 134778.Google Scholar
Cheng, H., Edwards, L.R., Chou Shen, C., Polyak, V.J., Asmerom, Y., Woodhead, J., Hellstrom, 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 371–372, 82e91.Google Scholar
Cisneros, M., Cacho, I., Frigola, J., Canals, M., Masqué, P., Martrat, B., Casado, M., et al., 2016. Sea surface temperature variability in the central-western Mediterranean Sea during the last 2700 years: a multi-proxy and multi-record approach. Climate of the Past 12, 849e869.Google Scholar
Cisneros, M., Cacho, I., Moreno, A., Stoll, H., Torner, J., Català, A., Edwards, R.L., Cheng, H., Fornós, J.J., 2021. Hydroclimate variability during the last 2700 years based on stalagmite multi-proxy records in the central-western Mediterranean. Quaternary Science Reviews 269, 107137.Google Scholar
Cnudde, V., Boone, M.N., 2013. High-resolution X-ray computed tomography in geosciences: a review of the current technology and applications. Earth-Science Reviews 123, 117.Google Scholar
Crowley, T.J., 2000. Causes of climate change over the past 1000 years. Science 289, 270277.Google Scholar
Crowley, T.J., Unterman, M.B., 2012. Technical details concerning development of a 1200 yr proxy index for global volcanism. Earth System Science Data 5, 187197.Google Scholar
Dandurand, G., Maire, R., Ortega, R., Devès, G., Lans, B., Morel, L., Perroux, A.S., et al., 2011. X-ray fluorescence microchemical analysis and autoradiography applied to cave deposits: speleothems, detrital rhythmites, ice and prehistoric paintings. Géomorphologie: relief, processus, environnement, Groupe français de géomorphologie 17, 407426.Google Scholar
Dasgupta, S., Saar, M.O., Edwards, R.L., Shen, C.C., Cheng, H., Alexander, E.C., 2010. Three thousand years of extreme rainfall events recorded in stalagmites from Spring Valley Caverns, Minnesota. Earth and Planetary Science Letters 300, 4654.Google Scholar
Debret, M., Sebag, D., Desmet, M., Balsam, W., Copard, Y., Mourier, B., Susperrigui, A.S., et al., 2011. Spectrocolorimetric interpretation of sedimentary dynamics: the new “Q7/4 diagram.” Earth-Science Reviews 109, 119.Google Scholar
Denniston, F., Luetscher, M.L., 2017. Speleothems as high-resolution paleoflood archives. Quaternary Science Reviews 170, 113.Google Scholar
Dorale, J.A., Lepley, S.W., Edwards, R.L., 2005. The ultimate flood recorder: flood deposited sediments preserved in stalagmites. Geophysical Research Abstracts 7, 09901.Google Scholar
Drobinsky, P., Da Silva, N., Bastin, S., Mailler, S., Muller, C., Ahrens, B., Christensen, O.B., Lionello, P., 2020. How warmer and drier will the Mediterranean region be at the end of the twenty-first century? Regional Environmental Change 20, 78.Google Scholar
Dumitru, O.A., Austermann, J., Polyak, V.J., Fornós, J.J., Asmerom, Y., Ginés, J., Ginés, A., Onac, B.P., 2019. Constraints on global mean sea level during Pliocene warmth. Nature 574, 233236.Google Scholar
Dumitru, O.A., Onac, B.P., Polyak, V.J., Wynn, J.G., Asmeron, Y., Fornós, J.J., 2018. Climate variability in the western Mediterranean between 121 and 67 ka derived from a Mallorcan speleothem record. Palaeogeography, Palaeoclimatology, Palaeoecology 506, 128138.Google Scholar
Edwards, R.L., Chen, J.H., Wasserburg, G.J., 1987. 238U-234U-230Th-232Th systematics and the precise measurement of time over the past 500,000 years. Earth and Planetary Science Letters 81, 175192.Google Scholar
Faust, J.C., Fabian, K., Milzer, G., Giraudeau, J., Knies, J., 2016. Norwegian fjord sediments reveal NAO related winter temperature and precipitation changes of the past 2800 years. Earth and Planetary Science Letters 435, 8493.Google Scholar
Finné, M., Kylander, M., Boyd, M., Sundqvist, H.S., Löwemark, L., 2015. Can XRF scanning of speleothems be used as a non-destructive method to identify paleoflood events in caves? International Journal of Speleology 44, 1723.Google Scholar
Fiol, L., Fornós, J.J., Gelabert, B., Guijarro, J.A., 2005. Dust rains in Mallorca (Western Mediterranean): their occurrence and role in some recent geological processes. Catena 63, 6484.Google Scholar
Fornós, J.J., Ginés, J., Gràcia, F., 2009. Present-day sedimentary facies in the coastal karst caves of Mallorca island (western Mediterranean). Journal of Cave and Karst Studies 71, 8699.Google Scholar
Frigola, J., Canals, M., Mata, P., 2015. Techniques for the non-destructive and continuous analysis of sediment cores. Application in the Iberian continental margin. Boletín Geológico y Minero 126, 609634.Google Scholar
Frisia, S., 2015. Microstratigraphic logging of calcite fabrics in speleothems as tool for palaeoclimate studies. International Journal of Speleology 44, 1.Google Scholar
Frisia, S., Badertscher, S., Borsato, A., Susini, J., Göktürk, O.M., Cheng, H., Edwards, R.L., Kramers, J., Tüysüz, O., Fleitmann, D., 2008. The use of stalagmite geochemistry to detect past volcanic eruptions and their environmental impacts. PAGES News 16, 2526.Google Scholar
Frisia, S., Borsato, A., Fairchild, I.J., Susini, J., 2005. Variations in atmospheric sulphate recorded in stalagmites by synchrotron micro-XRF and XANES analyses. Earth and Planetary Science Letters 235, 729740.Google Scholar
Frumkin, A., Stein, M., 2004. The Sahara–East Mediterranean dust and climate connection revealed by strontium and uranium isotopes in a Jerusalem speleothem. Earth and Planetary Science Letters 217, 451464.Google Scholar
Gao, C., Robock, A., Ammann, C., 2008. Volcanic forcing of climate over the past 1500 years: an improved ice core-based index for climate models. Journal of Geophysical Research 113, D23111.Google Scholar
Gázquez, F., Calaforra, J.M., Forti, P., Ghaleb, B., Delgado-Huertas, A., 2014. Paleoflood events recorded by speleothems in caves. Earth Surface and Landforms 39, 10, 13451353.Google Scholar
Gibelin, A.L., Deque, M., 2003. Anthropogenic climate change over the Mediterranean region simulated by a global variable resolution model. Climate Dynamics 20, 237339.Google Scholar
Ginés, J., Fornós, J., Ginés, A., Merino, A., Gràcia, F., 2014. Geologic constraints and speleogenesis of Cova des Pas de Vallgornera, a complex coastal cave from Mallorca Island (western Mediterranean). International Journal of Speleology 43, 2, 105124.Google Scholar
Giorgi, F., 2006. Climate change hot-spots. Geophysical Research Letters 33. doi:10.1029/2006GL025734.Google Scholar
Giorgi, F., Lionello, P., 2008. Climate change projections for the Mediterranean region. Global and Planetary Change 63, 90104.Google Scholar
González-Lemos, S., Müller, W., Pisonero, J., Cheng, H., Edwards, R.L., Stoll, H.M., 2015. Holocene flood frequency reconstruction from speleothems in northern Spain. Quaternary Science Reviews 127, 129140. https://doi.org/10.1016/j.quascirev.2015.06.002.Google Scholar
Goudie, A.S, Middleton, N.J., 2001. Saharan dust storms: nature and consequences. Earth-Science Reviews 56, 179e204.Google Scholar
Grimalt, M., Rosselló, J., 2018. Traditional flood mitigation measures in Mallorca. In: Antronico, L., Marincioni, F. (Eds.), Natural Hazards and Disaster Risk Reduction Policies. Geographies of the Anthropocene. II Sileno Edizioni, Lago, Italy, pp. 243260.Google Scholar
Grimalt-Gelabert, M., Bauzà-Llinás, J., Genovart-Rapado, M.C., 2021. The flood of October 9, 2018 in the city centre of Sant Llorenç des Cardassar (Mallorca). Cuadernos de Investigación Geográfica 47, 265286.Google Scholar
Grimalt-Gelabert, M., Rosselló-Geli, J., Bauzà-Llinàs, J., 2020. Flood related mortality in a touristic island: Mallorca (Balearic Island) 1960–2018. Journal of Flood Risk Management 13, e12644.Google Scholar
Hodge, E.J., 2004. Palaeoclimate of the Western Mediterranean Region: Results From Speleothems. Unpublished PhD thesis, University of Bristol, Bristol.Google Scholar
Hodge, E.J., Richards, D.A., Smart, P.L., Ginés, A., Mattey, D.P., 2008. Sub-millennial climate shifts in the western Mediterranean during the last glacial period recorded in a speleothem from Mallorca, Spain. Journal of Quaternary Science 23, 713718.Google Scholar
Homar, V., Ramis, C., Romero, R., Alonso, S., 2010. Recent trends in temperature and precipitation over the Balearic Islands (Spain). Climatic Change 98, 199211.Google Scholar
Jansen, J.H.F., Van der Gaast, S.J., Koster, B., Vaars, A.J., 1998. CORTEX, a shipboard XRF-scanner for element analyses in split sediment cores. Marine Geology 151, 143153.Google Scholar
Lionello, P., Malanotte-Rizzoli, P., Boscolo, R., 2006. Mediterranean Climate Variability. Elsevier, Amsterdam.Google Scholar
Lionello, P., Sanna, A., 2005. Mediterranean wave climate variability and its links with NAO and Indian monsoon. Climate Dynamics 25, 611623.Google Scholar
Lund, D.C., Lynch-Stieglitz, J., Curry, W.B., 2006. Gulf Stream density structure and transport during the past millennium. Nature 444, 601604.Google Scholar
Margaritelli, G., Cisneros, M., Cacho, I., Vallefuoco, M., Rettori, R., Lirer, F., 2018. Climatic variability over the last 3000 years in the central-western Mediterranean Sea (Menorca Basin) detected by planktonic foraminifera and stable isotope records. Global and Planetary Change 169, 179187.Google Scholar
Mariotti, A., Zeng, N., Yoon, J.H., Artale, V., Navarra, A., Alpert, P., Li, L.Z.X., 2008. Mediterranean water cycle changes: transition to drier 21st century conditions in observations and CMIP3 simulations. Environmental Research Letters 3, 044001.Google Scholar
Martín-Chivelet, J., Muñoz-García, M.B., Cruz, J.A., Ortega, A.I., Turrero, M.J., 2017. Speleothem architectural analysis: integrated approach for stalagmite-based paleoclimate research. Sedimentary Geology 353, 2845.Google Scholar
Martínez-Martínez, J., Fusi, N., Barberini, V., Cañaveras, J.C., Crosta, G.B., 2010. X-Ray Microtomography for studying 3D-textures of speleothems developed inside historic walls. Revista de la Sociedad Española de Mineralogía 13.Google Scholar
Martínez-Pillado, V.I., Iriarte, Y.E., Álvaro, A., Ortega, N., Aranburu, A., Arsuaga, J.L., 2020. The red coloration of Goikoetxe Cave's speleothems (Busturia, Spain): an indicator of paleoclimatic changes. Quaternary International 566–567, 141151.Google Scholar
Martín-Puertas, C., Jiménez-Espejo, F., Martínez-Ruiz, F., Nieto-Moreno, V., Rodrigo, M., Mata, M.P., Valero-Garcés, B.L., 2010. Late Holocene climate variability in the southwestern Mediterranean region: an integrated marine and terrestrial geochemical approach. Climate of the Past 6, 807816.Google Scholar
Mayewski, P.A., Maasch, K., Yan, Y., Kang, S., Meyerson, E., Sneed, S, Kaspari, S., et al., 2006. Solar forcing of the polar atmosphere. Annals of Glaciology 41, 147154.Google Scholar
McGregor, H.V., Evans, M.N., Goosse, H., Leduc, G., Martrat, B., Addison, J.A., Mortyn, P.G., et al., 2015. Robust global ocean cooling trend for the pre-industrial Common Era. Nature Geoscience 8, 671677.Google Scholar
Mees, F., Swennen, R., Van Geet, M., Jacobs, P., 2003. Applications of X-Ray Computed Tomography in the Geosciences. Geological Society, London.Google Scholar
Mickler, P.J., Ketcham, R.A., Colbert, M.W., Banner, J.L., 2004. Application of high-resolution X-ray computed tomography in determining the suitability of speleothems for use in paleoclimatic, paleohydrologic reconstructions. Journal of Cave and Karst Studies 66, 48.Google Scholar
Miller, G.H., Geirsdóttir, Á., Zhong, Y., Larsen, D.J., Otto-Bliesner, B.L., Holland, M.M., Bailey, D.A., et al., 2012. Abrupt onset of the Little Ice Age triggered by volcanism and sustained by sea-ice/ocean feedbacks. Geophysical Research Letters 39, L02708.Google Scholar
Mix, A.C., Rugh, W., Pisias, N.G, Veirs, S., 1992. Leg 138 Shipboard Sedimentologists (Hagelberg, T., Hovan, S., Kemp, a., Leinen, M., Levitan, M., Ravelo, C.), the Leg 138 Scientific Party, 1992. Color reflectance spectroscopy: a tool for rapid characterization of deep-sea sediments. Paper presented at the Proceedings of the Ocean Drilling Program. Part A, Initial Report 138, 6777. http://www.coas.oregonstate.edu/facultypages/Mix/Mix_etal_1992_ODP138ir_color.pdf.Google Scholar
Morellón, M., Valero-Garcés, B., González-Sampériz, P., Vegas-Vilarrúbia, T., Rubio, E., Rieradevall, M., Delgado-Huertas, , et al., 2011. Climate changes and human activities recorded in the sediments of lake Estanya (NE Spain) during the Medieval warm period and little Ice age. Journal of Paleolimnology 46, 423452.Google Scholar
Moreno, A., Cacho, I., Canals, M., Gromalt, J.O., Sanchez-Vidal, A., 2004. Millennial-scale variability in the productivity signal from the Alboran Sea record, Western Mediterranean Sea. Palaeogeography, Palaeoclimatology, Palaeoecology 211, 205219.Google Scholar
Moreno, A., Cacho, I., Canals, M., Prins, M.A., Sánchez-Goñi, M.F., Grimalt, J.O., Weltje, G.J., 2002. Saharan dust transport and high latitude glacial climatic variability: the Alboran Sea record. Quaternary Research 58, 318328.Google Scholar
Moreno, A., Valero-Garcés, B.L., González-Sampériz, P., Rico, M., 2008. Flood response to rainfall variability during the last 2000 years inferred from the Taravilla Lake record (Central Iberian Range, Spain). Journal of Paleolimnology 40, 943961.Google Scholar
Muhs, D.R., Budahn, J., Avila, A, Skipp, G., Freeman, J, Patterson, D., 2010. The role of African dust in the formation of Quaternary soils on Mallorca, Spain and implications for the genesis of Red Mediterranean soils. Quaternary Science Reviews 29, 25182540.Google Scholar
Muñoz, A., Bartolomé, M., Muñoz, A., Sancho, C., Moreno, A., Hellstrom, M.C.J., Osácar, C.M, Cacho, I., 2015. Solar influence and hydrological variability during the Holocene from a speleothem annual record (Molinos Cave, NE Spain). Terra Nova 27, 300311.Google Scholar
Muñoz-García, M.B., Cruz, J., Martín-Chivelet, J., Ortega, A.I., Turrero, M.J, López-Elorza, M., 2016. Comparison of speleothem fabrics and microstratigraphic stacking patterns in calcite stalagmites as indicators of paleoenvironmental change. Quaternary International 407, 7485.Google Scholar
Nederbragt, A.J, Thurow, J., 2004. Digital sediment colour analysis as a method to obtain high resolution climate proxy records. In: Francus, P. (Ed.), Image Analysis, Sediments and Paleoenvironments. Kluwer Academic, Dordrecht, Netherlands, pp. 105124.Google Scholar
Nieto-Moreno, V., Martínez-Ruiz, F., Giralt, S., Jiménez-Espejo, F., Gallego-Torres, D., Rodrigo-Gámiz, M., García-Orellana, J., Ortega-Huertas, M., de Lange, G. J., 2011. Tracking climate variability in the western Mediterranean during the Late Holocene: a multiproxy approach. Climate of the Past 7, 13951414.Google Scholar
Parnell, A.C., Haslett, J., Allen, J.R.M., Buck, C.E, Huntley, B., 2008. A flexible approach to assessing synchroneity of past events using Bayesian reconstructions of sedimentation history. Quaternary Science Reviews 27, 18721885.Google Scholar
Pastor, F., Estrela, M.J., Peñarrocha, D., Millán, M.M., 2001. Torrential rains on the Spanish Mediterranean coast: modelling the effects of the sea surface temperature. Journal of Applied Meteorology 40, 11801195.Google Scholar
Pino, D., Alberola, A., Balasch, J.C., Barriendos, M., Gil, S., Grau-Satorras, M., Mazón, J., Pérez Morales, A., Tuset, J., 2018. Major flood events reconstruction from a multi-proxy approach. The case study of November 1617 flood event in the Mediterranean Basins of Iberian Peninsula. Geophysical Research Abstracts 20, EGU2018-10386.Google Scholar
Polyak, V.P., Onac, B.P., Fornós, J.J., Hay, C., Asmeron, Y., Dorale, J.A., 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
Ramos-Román, M.J., Jiménez-Moreno, G., Camuera, J., García-Alix, A., Anderson, R.S., Jiménez-Espejo, F.J., Carrión, J.S., 2018. Holocene climate aridification trend and human impact interrupted by millennial- and centennial-scale climate fluctuations from a new sedimentary record from Padul (Sierra Nevada, southern Iberia Peninsula). Climate of the Past 14, 117137.Google Scholar
Robock, A., 2000. Volcanic eruptions and climate. Reviews of Geophysics 38, 191219.Google Scholar
Rodrigo-Gámiz, M., Martínez-Ruiz, F., Jiménez-Espejo, F.J., Gallego-Torres, D., Nieto-Moreno, V., Romero, O., Ariztegui, D., 2011. Impact of climate variability in the western Mediterranean during the last 20,000 years: oceanic and atmospheric responses. Quaternary Science Reviews 30, 20182034.Google Scholar
Rogerson, M., Weaver, P.P.E., Rohling, E.J., Lourens, L.J., Murray, J.W, Hayes, A., 2006. Color logging as a tool in high-resolution paleoceanography. In: Rothwell, R.G. (Ed.), New Techniques in Sediment Core Analysis. Geological Society of London Special Publication 267, 99112.Google Scholar
Rothwell, R.G, Rack, F.R., 2006. New techniques in sediment core analysis: an introduction. In: Rothwell, L.G. (ed.), New Techniques in Sediment Core Analysis. Geological Society of London Special Publication 267, 129.Google Scholar
Sheffield, J., Wood, E.F., 2008 Projected changes in drought occurrence under future global warming from multi-model, multi scenario, IPCC AR4 simulations. Climate Dynamics 31, 79105.Google Scholar
Thorndycraft, V.R., Barriendos, M., Benito, G., Rico, M., Casas, A., 2006. The catastrophic floods of AD 1617 in Catalonia (northeast Spain) and their climatic context. Hydrological Sciences Journal 51, 5, 899912.Google Scholar
Torner, J., Cacho, I., Moreno, A., Sierro, F.J., Martrat, B., Rodriguez-Lazaro, J., Frigola, J., et al., 2019. Ocean–atmosphere interconnections from the last interglacial to the early glacial: an integration of marine and cave records in the Iberian region. Quaternary Science Reviews 226, 106037.Google Scholar
Tsimplis, M.N., Josey, S.A., 2001. Forcing of the Mediterranean Sea by atmospheric oscillations over the North Atlantic. Geophysical Research Letters 28, 803806.Google Scholar
Ulbrich, U., May, W., Li, L., Lionello, P., Pinto, J. G., Somot, S., 2006. The Mediterranean climate change under global warming. In: Lionello, P., Malanotte-Rizzoli, P., Boscolo, R. (Eds.), Mediterranean Climate Variability. Elsevier, Amsterdam, pp. 398415.Google Scholar
Vanghi, V., Iriarte, E., Aranburu, A., 2015. High resolution X-ray computed tomography for petrological characterization of speleothems. Journal of Cave and Karst Studies 77, 7582.Google Scholar
Vansteenberge, S., Winter, N.J., Sinnesael, M., Xueqin, Z., Verheyden, S., Claeys, P., 2020. Benchtop μXRF as a tool for speleothem trace elemental analysis: Validation, limitations and application on an Eemian to early Weichselian (125–97 ka) stalagmite from Belgium. Palaeogeography, Palaeoclimatology, Palaeoecology 538, 109460.Google Scholar
Vesica, P.L., Tuccimei, P., Turi, B., Fornós, J.J., Ginés, A, Ginés, J., 2001. 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
Walczak, I.W., Baldini, J.L., Baldini, L.M., McDermott, F., Marsden, S., Standish, C.D., Richards, D.A., Andreo, B., 2015. Reconstructing high-resolution climate using CT scanning of unsectioned stalagmites: a case study identifying the mid-Holocene onset of the Mediterranean climate in southern Iberia. Quaternary Science Reviews 127, 117128.Google Scholar
Westland, S., 2012. Frequently Asked Questions About Colour Physics. Kindle Edition.Google Scholar
Zisu, N.S., Schwarcz, H.P., Konyer, N., Chow, T., Noseworthy, M.D., 2012. Macroholes in stalagmites and the search for lost water. Journal of Geophysical Research: Earth Surface 117. doi:10.1029/2011JF002288.Google Scholar
Figure 0

Figure 1. Studied area. (a) Map of the western Mediterranean showing Balearic Islands, where Mallorca is located. (b) Distribution of the superficial hydrology of Mallorca Island; streams are indicated in blue and watersheds in red (adapted from Grimalt-Gelabert et al., 2020). Location of Sa Balma des Quartó cave is indicated as well as the municipalities most affected by the severe flood that occurred on October 9, 2018 (Grimalt-Gelabert et al., 2021). (c) Topography of Sa Balma des Quartó cave (Bermejo et al., 2014). Position of the vertical profile in part d is also indicated. Blue circles correspond to the location of the speleothems recovered in situ. (d) Vertical profile (“G-g” section in Bermejo et al., 2014). Filled circles in c and d represent sites where evidence of cave flooding was observed after the extreme rain event of October 2018.

Figure 1

Figure 2. Pictures of evidence of flood into Sa Balma des Quartó cave in November 2018: (a and b) in the upper part of the cave, gours containing water (blue lines); (c) in the lower part of the cave, marks of water flows (purple lines).

Figure 2

Figure 3. Synthetized description of the Seán stalagmite (from Cisneros et al., 2021). The brown dashed line indicates the layer/unconformity ca. 6.00 cm. Shown in relation to the fabric pictures are those around 6.00 cm (from top to bottom): top, isolated mosaic fabrics (Mo) above the micrite layer (m), columnar fabric below (C), cross-polarized light; bottom, columnar (C), mosaic (Mo), and micrite (m) fabrics, plane-polarized light. U-Th age model performed on Bchron. Red plots are the final age models. Black diamonds represent the U-Th ages (2σ error). Grey dashed lines correspond to the total range of ages covered by all the age models obtained with Bchron, which have statistically the same significance (95&per; confidence interval).

Figure 3

Table 1. Summary of the results of U-Th analyses used in the age model of the Seán stalagmite (2σ error) and published in Cisneros et al. (2021).

Figure 4

Figure 4. Fabric picture of the Seán stalagmite thin section ca. 6.00 cm. Columnar (C), mosaic (Mo), and micrite (m) fabrics, plane-polarized light. The arrow indicates the gastropod test.

Figure 5

Figure 5. Micro-computed tomography (micro-CT) and X-ray fluorescence (XRF) core scanning (XRFCS) results (this study) from the boundary of the two unconformity-bounded units (UBUs) in Seán (4.75–7.75 cm). Left, Colour map from micro-CT between (HU, Hounsfield units); horizontal grey bar indicates the detrital layer (from 6.00 cm to 6.25 cm). Right panel (from left to right), Coordinates of colour CIE L (Lightness)*a*b*; S, Ti, Ca, and Mn elements (expressed as peak area). XRFCS measurements were carried out on the left part of the growth axis of the stalagmite. Vertical and discontinuous black line (above the left image) indicates the line of measurements.

Figure 6

Figure 6. Results presented in this study from the Seán stalagmite compared with previously published data for the Seán and Multieix stalagmites, North Atlantic Oscillation (NAO) reconstruction and volcanism activity. (a–e) Ba/Ca, Sr/Ca, Mg/Ca, δ13C, and δ18O records for the Seán stalagmite. δ13C, δ18O records from Multieix are also shown in lighter colours (Cisneros et al., 2021). (f and g) S and Ti from X-ray fluorescence (XRF) core scanning (XRFCS) analyses (this study). (h and i) Growth rates and U/Th ages (diamonds) of Seán and Multieix stalagmites (Cisneros et al., 2021). (j) Mg/Ca sea-surface temperature (SST) from north Minorca (Cisneros et al., 2016). (k) NAO reconstruction (Faust et al., 2016). (l) Northern and global volcanism (Gao et al., 2008; Crowley and Unterman, 2012). Both subperiods of Little Ice Age (LIAa and LIAb) and the Industrial Era (IE) are also indicated. Brown vertical band indicates discontinuity and brown layer observed in Seán (~6.00 cm), which corresponds to the limit between both unconformity-bounded units (UBUs) in the Seán stalagmite.

Figure 7

Figure 7. Hypothetical reconstruction of the cave flooding during the seventeenth century. Topography corresponds to the vertical profile of Sa Balma des Quartó cave (“G-g” section in Bermejo et al., 2014) also shown in Fig. 1d. Blue circles correspond to the locations where the speleothems were found, which are separated by a distance of ~5 m. Seán's growth position is located approximately 2 m lower than the Multieix site. The hypothetical flood level, which is represented in qualitative terms, would have covered the 6.00 cm length of the Seán stalagmite in that moment. The evidence of cave flooding observed after the extreme rain event of October 2018 (gours with water in the upper part and water flow marks in the lower part) is also indicated.