dedication - university of south florida

119
Assessing Paleo-Coastal Caves Evolution and Sea Level Changes in Mallorca: Results from Radiometric Dating of Speleothems by Giuseppe Lucia A thesis submitted in partial fulfillment of the requirements for the degree of Master of Science in Geology School of Geosciences College of Arts and Sciences University of South Florida Major Professor: Bogdan P. Onac, Ph.D. Joan J. Fornós, Ph.D. Victor J. Polyak, Ph.D. Ping Wang, Ph.D. Date of Approval: March 19, 2019 Keywords: Karst Geomorphology, Paleokarst, Flowstones, Stalagmites, U-series Copyright © 2019, Giuseppe Lucia

Upload: others

Post on 11-Dec-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Assessing Paleo-Coastal Caves Evolution and Sea Level Changes in Mallorca:

Results from Radiometric Dating of Speleothems

by

Giuseppe Lucia

A thesis submitted in partial fulfillment

of the requirements for the degree of

Master of Science in Geology

School of Geosciences

College of Arts and Sciences

University of South Florida

Major Professor: Bogdan P. Onac, Ph.D.

Joan J. Fornós, Ph.D.

Victor J. Polyak, Ph.D.

Ping Wang, Ph.D.

Date of Approval:

March 19, 2019

Keywords: Karst Geomorphology, Paleokarst, Flowstones, Stalagmites, U-series

Copyright © 2019, Giuseppe Lucia

DEDICATION

To my family, for the love and support essential in my life. To the ‘Free World Riders’,

because without them I could not be the person I am. Love you all.

ACKNOWLEDGEMENTS

The research was funded by a collaborative National Science Foundation grant AGS

1602670 to Bogdan P. Onac and Victor J. Polyak.

I am deeply grateful to Bogdan for his guidance and support over these two years. Special

thanks to Victor for his patience and crucial help during the U-series analyses and to the

improvements of the manuscript. Joan J. Fornós and Joaquín Ginés are acknowledged for their

logistic and assistance during sampling campaigns and the suggestions to enhance the thesis. I

would also like to thank my colleague Oana Dumitru for her help with the lab work and

encouragement over the last period of writing. I am also thankful to Viorel Atudorei for his support

during my stay in Albuquerque, NM. Thanks to my colleagues Oana, Robert, Abby, Daniel,

Dylan, Isidoros, and my roommates Renato and Hari for their moral support and the great time

spent together over these two years.

I would like to thank my family and friends in Italy and my American relatives for their

love and help that is extremely important to me. Finally, I am grateful to Jo De Waele who

introduced me to the fascinating underground realm.

i

TABLE OF CONTENTS

List of Tables ................................................................................................................................. iii

List of Figures ................................................................................................................................ iv

Abstract ............................................................................................................................................x

Chapter One: Introduction ...............................................................................................................1

1.1 Aim of the Research.......................................................................................................3

1.2 Thesis Structure .............................................................................................................3

Chapter Two: Background ...............................................................................................................4

2.1 Study Area Description ..................................................................................................4

2.1.1 Geographic and Geomorphologic Settings .....................................................4

2.1.2 Climate ............................................................................................................6

2.1.3 Geologic Settings ...........................................................................................6

2.1.4 Mallorcan Karst: Speleogenetic Processes and Cave Types...........................8

2.1.5 Paleokarst Features: Structural and Morphogenetic Aspects .............................. 11

2.2 U-Th-Pb Geochronology of Speleothems ....................................................................15

2.2.1 U, Th, and Pb in Natural Environments ........................................................15

2.2.2 Overview of Radioactive Decay Mechanisms and Laws .............................16

2.2.3 U-Th-Pb Decay Systems and Dating Methods .............................................18

2.2.4 U-Pb Age Determination Methods ........................................................... 20

2.2.5 Previous U-Pb Dating Studies of Speleothems ......................................... 21

2.2.6 U-series Age Determination ..................................................................... 24

2.2.7 Previous U-series Investigations of Secondary Cave Deposits ....................25

2.3 Oxygen and Carbon Isotope Geochemistry of Speleothems .......................................25

Chapter Three: Materials and Methods ..........................................................................................27

3.1 Investigation Sites and Samples Description ...............................................................27

3.2 U-series and U-Pb Dating Analyses ..................................................................................... 35

3.2.1 Samples Preparation............................................................................................... 35

3.2.2 Chemical Procedures ............................................................................................. 35

3.2.3 Analytical Methods .............................................................................. 37

3.2.4 Data Processing, Age Models and Calculations ........................................ 38

3.3 Oxygen and Carbon Isotopic Analysis ........................................................................39

ii

Chapter Four: Results ....................................................................................................................41

4.1 Cala Falcó ............................................................................................................................... 41

4.1.1 Stalagmite CF-11 ...................................................................................... 43 4.2 Dimoni .................................................................................................................................... 46 4.3 Sa Ferradura ............................................................................................................................ 47

4.4 Na Mitjana ........................................................................................................................ 48 4.4.1 Sample MIP-5 ........................................................................................... 50

4.5 Coves del Drac ........................................................................................................ 51 4.5.1 Speleothem DR-D20 ................................................................................. 51

4.6 U-series Systematics and Ages of Investigated Speleothems ......................................52

Chapter Five: Discussion ...............................................................................................................55

5.1 Mallorcan Coastal Karst and Relative Sea Level Changes over the Past

750 kyr .........................................................................................................................55

5.1.1 Cala Falcó Paleo-cave ...................................................................................62

5.1.2 Sa Ferradura and Dimoni Paleokarst ............................................................64

5.1.3 Na Mitjana Paleokarst ...................................................................................67

5.2 Constraining Relative Sea Level Maximum in Mallorca during the MIS 7 ................69

5.3 MIS 5b/5a Transition and MIS 5a Sea Level Highstand in Mallorca..........................73

Chapter Six: Conclusions ...............................................................................................................79

References ......................................................................................................................................82

Appendix A: Data ..........................................................................................................................96

Appendix B: Copyright Permissions ...........................................................................................103

iii

LIST OF TABLES

Table A1. U-series dataset and ages of the samples in this study.................................................97

Table A2. U-Pb data for the flowstone FE-1-A ............................................................................98

Table A3. Carbon and Oxygen isotopic data for CF-11 stalagmite ..............................................99

Table A4. Age-depth model dataset for CF-11 stalagmite generated from StalAge

Algorithm (Scholz & Hoffmann, 2011) .......................................................................101

iv

LIST OF FIGURES

Figure 1. Localization of Mallorca Island in the western Mediterranean. It is visible the

structural alignment NE-SW of the Balearic Promontory and the surrounding

basins. Image modified from Google Earth ....................................................................5

Figure 2. Simplified geological map of Mallorca showing the principal morphostructural

and stratigraphic units (modified from Fornós et al., 2019) ...........................................7

Figure 3. Idealized model of the different cave typologies developing in the island of

Mallorca (modified from Ginés and Ginés, 2011) ........................................................10

Figure 4. A) Schematic summary of the paleokarst development in the Upper Miocene

limestones (modified from Fornós, 1999). Syn-sedimentary paleokarst

structures: B) The eastern coastal zone of Portocolom village, and C) Punta

des Savinar (Mondragó). Post-sedimentary paleokarst features: D) Littoral

area near the entrance of Cova del Dimoni (Porto Cristo), and E) Cova des

Pont (Manacor) (photo credit: B.P. Onac) ............................................................ 13

Figure 5. Morphological evolution of paleokarst structures in the eastern coastal region

of Mallorca (Solimina, Mondragó Natural Park) (photo credit: J.G. Ginés)

(see text for explanations) .............................................................................................14

Figure 6. Uranium (green), Actinium (yellow), and Thorium (orange) decay series.

Intermediate radioactive daughter and half-lives (>10 years) are shown.

Decay constants of 238U, 235U, and 232Th are dispayed (from Schoene, 2014)..............18

Figure 7. Map of Mallorca showing the location of the sampling sites: Cala Falcó (CF),

Coves del Drac (DR), Dimoni (DIP), sa Ferradura (FE), and na Mitjana (MIP).

All locations are color coded.........................................................................................27

Figure 8. Cala Falcó site: A) Overview of the cala and the cave entrance; B) Panorama of

the paleokarst structure outcropping on the cliff; C-E) Close up of the sampling

points (photo credits: B.P. Onac & J. McIlrath) ...........................................................28

v

Figure 9. A) Dimoni outcrop the breccia deposits with flowstones and stalagmites exposed,

both truncated and/or embedded; B) Close up of the red polymictic breccias;

C) DIP-1-A and D) DIP-1-B flowstones, respectively (photo credits: B.P. Onac

& J. McIlrath) ........................................................................................................... 30

Figure 10. The 2 m thick section of flowstone located in the sa Ferradura site. The

sampling points are shown: FE-1-A (lower), FE-1-E (middle), and FE-1-F

(top) (photo credits: B.P. Onac) ....................................................................................31

Figure 11. Na Mitjana site. A: Cemented blocks in which embedded speleothems are

exposed; B: MIP-1 and MIP-2 sampling points; C: Small cavity on the right

side of na Mitjana Cave entrance; D: Close up of MIP-3 sampling point;

E: Close up of the flowstone from which MIP-4-A and MIP-4-B were collected;

F: Photo of the MIP-5 sample (photo credits: B.P. Onac) ............................................32

Figure 12. A) Aerial picture of the Portocristo area and location of the Coves del Drac

(from Ginés et al., 2014), B) Image of the Llac Delícìes (photo credit: J. Ginés),

C) Stalagmite DR-D20 showing in its outer surface weathered white layers;

and D) Sample cross-section displaying soda straw, stalactite, and stalagmite

components with the thin white weathered laminae on the upper part .................. 34

Figure 13. Sub-samples preparation and chemical proceedings. Sub-samples are

crushed (A), and stored in tubes (B); C) Dissolved aliquots are spiked,

purified and dried down on hotplate; D-E) U, Th, and Pb are chemically

separated (see text for detailed explanation) ........................................................ 36

Figure 14. Cup configurations and resistors settings on the Thermo NEPTUNE

MC-ICPMS for the U (A), Th (B), and Pb (C) isotopic measurements

(modified from LaPointe, 2013) .......................................................................... 38

Figure 15. Speleothem samples from Cala Falcó used for U-series dating. The yellow

bands mark the position and code of each subsample...................................................42

Figure 16. Stalagmite CF-11. A) Cross section of the sample showing its individual growth

layers and internal corrosion features. The subsamples (color bands) used for

U-series dating are also marked. CF-11-21 (grey band) denotes the outlier age.

The double white lines represent brief hiatuses (corrosion layers) at 20 and 11 mm

from the top, respectively. B) Age-depth model for CF-11 stalagmite generated

from StalAge algorithm plotting the MC-ICP-MS U-series ages against distance

from top .................................................................................................................................... 43

vi

Figure 17. Curve plots 232Th vs. Initial 230Th/232Th correction for all the subsamples

(color-coded) and their corresponding absolute 2σ errors. The equation

used to calculate the initial 230Th/232Th correction is also shown. The

equation used to calculate the initial 230Th/232Th correction is also displayed.

The x values were assumed for each subsample from different 230Th

concentrations in accordance with the stratigraphic order ........................................44

Figure.18. Records of δ18O and δ13C values for the stalagmite CF-11 ..........................................45

Figure 19. Dimoni flowstones: A) DIP-1A-1 and B) DIP-1B. Subsamples’ code and

position are also displayed (yellow marks) ...................................................................47

Figure 20. Flowstone samples from sa Ferradura site: A) FE-1-A; B) FE-1-E; and

C) FE-1-F. Subsamples location (in yellow) and related codes are also shown ...........48

Figure 21. Na Mitjana samples: A) MIP-1; B) MIP-2; C) MIP-3; D) MIP-4A; and E) MIP-4B.

Subsamples position (in yellow) and code are indicated for each speleothem ................... 49

Figure 22. MIP-5 sample cross section. Powder subsamples and codes are displayed

in yellow. Hiatuses at 20 and 23 mm (from the outer surface) are shown

in white .........................................................................................................................50

Figure 23. Cross section of sample DR-D20. Morphological components of the speleothem

are indicated and labeled. The original soda straw is shown in orange (dashed)

and the stalactite-stalagmite contact in black (line). Yellow: subsamples for

U-series ages used in this study; Grey: subsamples that provided unreliable

U-series ages; Red: subsamples of whose ages were discarded and not included

in the study ....................................................................................................................52

Figure 24. Uranium-evolution curve for the speleothems in this study. All samples are

color-coded and described in the legend. In close-up, the samples clustering near

the secular equilibrium line. Ellipse dimensions express absolute 2σ errors ................53

Figure 25. Plot of all ages from all speleothems investigated in this study. Samples are

site-based color-coded and labeled. Errors bars are absolute 2σ. Shaded area

represents the U-series limit range ................................................................................54

vii

Figure 26. Ages of the speleothems from our study and eustatic sea level (ESL)

reconstructions from Naish and Wilson (2008) (orange curve), and Spratt and

Lisiecki (2016) (blue curve). Speleothem samples are color-coded. All age

errors are absolute 2σ. Gray shaded box indicates the period of possible

breakdown and dismantling of the paleo-caves investigated in this study.

Marine Isotopes Stages (MIS) are also labeled ............................................................57

Figure 27. Conceptual model showing the different speleogenetic stages in the evolution

of the coastal caves in the eastern Mallorca [modified after Ginés and Ginés

(1992) and Ginés et al. (2014)]: 1) Cave genesis in the mixing zone, 2) Sea level

rise and further cave development, 3) Collapses following sea level drop,

4) Vadose speleothems precipitation, 5) Partial collapse of ceiling and

cementation of breakdowns, and 6) Coastal erosion that exposes speleothems at

surface. See text for detailed explanation .....................................................................59

Figure 28. Images of Cala Falcó site: A) General view of the cala, cave entrance, and

collapsed blocks, B) Image of the paleokarst features outcropping on the cliff

outside the cave, and C) Close up of the paleo-flowstones and stalagmites.

Yellow stars represent the sampling points, whereas the white lines highlight

the contour of the speleothems (photo credits: B.P. Onac and J. McIlrath) .................63

Figure 29. Sa Ferradura and Dimoni sites: A) Aerial view of the two engulfments,

B) Panoramic photograph of the Sa Ferradura site, C) Flowstone sequence

outcropping in between breccia deposits, D) Fossil flowstones and stalagmites

exposed at Dimoni, and E) Funnel-like embayment produced by the Messinian

paleokarst structure and cave developed in the breccia at present sea level

(photo credits: B.P. Onac, J.J. Fornós, and J. Ginés) ....................................................66

Figure 30. Na Mitjana site: A) Paleokarst structures and cavities (highlighted), B) Various

speleothems in one of the cavities, C) Close image of the sample MIP-5,

D) Mammillary-like speleothem MIP-3, and E) Flowstone MIP-4 (photo

credits: B.P. Onac) ........................................................................................................68

Figure 31. Image of Llac de les Delícies (Coves del Drac) showing the hypothetical phases

of DR-D20 precipitation: a) soda straws/stalactites reaching the cave floor,

b) hanging portions of broken stalactites c) stalagmite covering the broken part

of a soda straw/stalactite at the cave floor (photo credits: J.J. Fornós) .........................71

viii

Figure 32. MC-ICP-MS U-series ages for DR-D20 speleothem (red triangle) are

compared with the Argentarola stalagmites (Bard et al., 2002; Dutton

et al., 2009) (orange, purple and black squares), POS TIMS and α-spectrometry

ages from Grotta del Nettuno, Sardinia (Tuccimei et al., 2000a), and POS α-

spectrometry ages from Cove den Passol and Na Mitjana Cave, Mallorca

(Tuccimei et al., 2000b; Vesica et al., 2000). RSL curve (Waelbroeck et al.,

2002) and Red Sea level reconstruction (Siddall et al., 2003) are also shown

(solid and dashed blue lines). All errors are absolute 2σ, except for α-spectrometry

ages from Tuccimei et al. (2000a, b), which are 1σ) ....................................................72

Figure 33. Conceptual drawing of CF-11 stalagmite evolution and sea level implications:

A) stalagmite precipitation in vadose conditions during sea level lowstand,

B) cave partial flooding during MIS 5a sea level higstand cause the cessation of

the stalagmite growth, and C) sea level fall and subaerially exposure of the

stalagmite ......................................................................................................................75

Figure 34. Comparison of Cala Falcó (CF-11) stalagmite oxygen composition with other

local and global records: a) NGRIP (North Greenland Ice Core Project members,

2004; Rasmussen et al., 2014), b) δ18O records of CF-11, c) Campanet stalagmite

δ18O values (Dumitru et al., 2018), d) δ18O of Cala Falcó (CFF-03-03-1)

stalagmite (Hodge et al., 2008), and e) Sea level reconstructions for Mallorca

during MIS 5a, based on phreatic overgrowths on speleothems (Tuccimei

et al., 2006; Dorale et al., 2010) (blue and red circles, respectively). U-series

ages of the POS (2σ absolute errors) and elevations related to present sea level

are shown. MIS boundaries are also displayed. MIS 5a highstand is shown in

light blue .......................................................................................................................76

Figure 35. δ13C comparison between: a) Cala Falcó (CFF-03-03-1) stalagmite (Hodge et

al., 2008), b) Cala Falcó (CF-11) stalagmite, and c) Campanet speleothem

(Dumitru et al., 2018), MIS 5b/5a boundary is shown..................................................78

Figure B1. Authorship permission provided by Joaquin Ginés for using a picture from

Mondragó Natural Park in the present study (Fig. 5)..................................................103

Figure B2. Permission granted from Joaquin Ginés to use his images from Mallorca

Island in the present study (Figs. 12 and 29) ..............................................................103

Figure B3. Permission provided from Joan Fornós to use some pictures in this thesis

(Fig. 29 and 31 ............................................................................................................104

ix

Figure B4. Permission granted by Judy Mcllrath to use some pictures of Mallorca

Island in this study (Figs. 8A, 9B-D, and 29A) ..........................................................104

Figure B5. Permission granted by Bogdan P. Onac to use the pictures of his property

in the present study (Figs. 4, 8, 9, 10, 11, 28, 29, and 30) ..........................................105

Figure B6. Permission obtained by Elsevier to use a picture (Fig. 6 in the present

study) originally published in the book Treatise on Geochemistry ............................106

x

ABSTRACT

Sea level affects the littoral morphology and structure in different ways. In coastal karst basins,

dissolutional and depositional processes are strictly related to sea level variation. A great variety of

karst-related geomorphological features and cave deposits exist above, below, and at sea level, and are

extremely useful to study past changes of relative sea level. Within these, vadose speleothems such as

flowstones and stalagmite, can document unequivocal maximum sea level constraints as well as record

phreatic phases of sea level rise. Here we present a series of geochronological data from the eastern

Mallorca and show the implication of relative sea level oscillations in the coastal karst evolution over

the last 750 kyr. We investigate a total number of twenty-two samples between flowstones and

stalagmites from different paleokarst sites and coastal caves. U-series ages of samples ranged between

650 and 450 kyr for the paleo-flowstones, and 225-198 ka and 85.7-83.7 ka for the stalagmites. The

paleo-speleothems records provide new timing constraints regarding periods of vadose conditions

suggesting that sea level did not exceed 2 mapsl during the interglacial highstands 17, 15, and 13.

Periodic sea level oscillations following glacial/interglacial cycles caused morphostructural changes of

the paleo-caves. We constrain the occurrence of breakdown events to intervals succeeding the sea

level highstands 11, 9, 7, and 5, and further total breakdowns and dismantling of caves throughout the

Holocene. We also present two stalagmite records of relative sea level maximum at ~1 mbpsl during

the MIS 7, and constraints for sea level rise and paleoenvironmental changes during the MIS 5a. We

conclude pointing out the necessity to use vadose speleothems along with other RSL records to

precisely estimate past sea level changes.

1

CHAPTER ONE:

INTRODUCTION

Sea level change is among the most debated and studied research topic in science,

because of its direct concern to human society and economy (Clark et al., 2016).

Understanding the mechanisms controlling sea level fluctuations, such as the connection

between polar-ice sheet growth/melt and changing global temperatures, help us to better

estimate and predict potential future scenarios (Dutton and Lambeck, 2012; DeConto and

Pollard, 2016). Nevertheless, data sets generated from accurate monitoring and direct

observations are not sufficient per se to produce future projections, and the study of past sea

level oscillations is somehow essential to accomplish this purpose (Dutton et al., 2015).

Variations in mass and /or volume of ocean water in response to ice-sheet growth/

reduction is reflected in global changes of mean sea level and is defined as eustatic sea level

(ESL) (Suess, 1906; Gornitz, 2005), whereas variations of sea level related to any land-linked

reference point is defined as relative sea level (RSL) (Kemp et al., 2015). RSL is mainly

associated with land movement (e.g., uplift, subsidence, etc.) and may or may not directly

reflect global mean sea level changes (Rovere et al., 2016). However, many locations

worldwide present moderately stable tectonic settings and host coastal geologic archives that

are extremely useful to investigate past sea level variations.

In the western Mediterranean region, Mallorca represents a benchmark site for paleo

sea level studies, due to its steady, low tectonic movements and the great variety of coastal

records and karst-based sea level indicators that recorded sea level fluctuations over the past

2

~10 million years (Fornós et al., 2002a; Giménez and Gelabert, 2002; Just et al., 2011; Ginés et

al., 2012). Coastal geomorphologic RSL archives include beach rocks, marine terraces, tidal

notches, shore platforms, coral reef terraces, beach deposits, and fossil fixed biological

indicators (Rovere et al., 2016). The structural and morphological features of littoral records

contributed to the estimate the relative sea level elevations during the warm intervals MIS 5e

and MIS 11 in Mallorca (Cuerda, 1989; Zazo et al., 2003, 2013; Lorscheid et al., 2017).

Among the coastal karst deposits having a great importance for sea-level studies are the

phreatic overgrowths on speleothems (POS), defined as carbonatic encrustations that

precipitate over preexisting speleothems and cave walls in brackish water sea level related

pools (Pomar et al., 1979; Ginés, 2000). Previous investigations of POS in Mallorca have

precisely documented paleo sea level positions during the Holocene (Tuccimei et al., 2010),

the last interglacial period (Vesica et al., 2000; Tuccimei et al., 2006; Dorale et al., 2010;

Polyak et al., 2018), as well as throughout the middle Pliocene (Dumitru et al., 2018a). Beside

these, vadose speleothems such as stalagmites, flowstones, and stalactites, precipitated in

coastal caves also provide indirect information about sea level, particularly, they define the

maximum sea level boundary during periods of relatively low sea-stands, as well as indicate

phases of relative sea level rise that caused floods in the caves stopping the precipitation of

carbonate speleothems (Onac et al., 2012).

A series of radiometric data on stalagmites and flowstones from different paleokarst structures

and coastal caves of Mallorca is presented. These results will be used to discuss the

correlation between coastal karst processes and sea level variations that occurred on this island

over the past 750 kyr. The study also places minimum and maximum age constraints for the

timing of sea-level rise and fall during the penultimate and the last interglacial periods.

3

1.1 Aim of the Research

Our project seeks to improve the knowledge regarding coastal dynamics, karst processes, and

sea level changes in Mallorca. By analyzing the vadose speleothems exposed in coastal paleokarst

structures along the eastern coast of the island, an attempt will be made to provide new timing

constraints in the speleogenetic phases of the paleo-coastal caves. Another goal is to precisely

describe the periods when vadose conditions existed that allowed the precipitation of thick sequences

of flowstones and stalagmites, at times when sea level regressed. The use of new radiometric data

will be used to offer the previous coastal cave genetic models a geochronological timeframe, and to

define the intervals when the breakdown of the paleo-caves occurred. Furthermore, U-series dating

and oxygen and carbon stable isotope investigations on stalagmites from two littoral caves were used

to characterize the paleoenvironmental conditions and sea level variations during the past

glacial/interglacial cycles in the western Mediterranean.

1.2 Thesis Structure

The thesis consists of six chapters. Chapter One will introduce the reader to the topics

described in the research. Various information regarding the geological, geomorphological, and

speleological characteristics of Mallorca, along with some background data regarding the analytical

techniques are given in Chapter Two. Detailed descriptions of the investigated sites and samples,

along with the methodology, are described in Chapter Three. The results of the U-series, U-Pb, and

stable isotope geochemical analyses are presented in Chapter Four. In Chapter Five I interpret the

radiometric age data of various vadose speleothems in order to add new chronological insights into

the speleogenetic phases and coastal karst evolution, both directly related to changes in sea level. I

will also provide new karst-based sea level constraints during the penultimate and ultimate

interglacials. The thesis ends with a summary of outcomes and conclusions in Chapter Six.

4

CHAPTER TWO:

BACKGROUND

2.1 Study Area Description

2.1.1 Geographic and Geomorphologic Settings

Mallorca Island extends between the 39°58’ N (Cap de Formentor), 3°28’ E (Punta de

Capdepera), 39°15’ N (Cap de ses Salines), and 2°17’ E (Sant Elm), with a NE-SW

arrangement of its major structural geomorphic units. The island represents the widest

emerged portion of the Balearic Promontory, with a surface area of ~3,650 km2 and a perimeter

of about 560 km. The promontory is surrounded by the Valencia Basin (NW), the Liguro-

Provencal Basin (NE), the Algerian Basin (SSE) and a steep escarpment (Emile Baudot

Escarpment) on the SE side (Fig. 1). Geomorphologically, Mallorca can be subdivided into

three morphostructural units: Serra de Tramuntana, Es Pla, and Serres de Llevant, all of which

are the result of complex extensional features of horsts and grabens (Gelabert, 1998). Serra de

Tramuntana is the most remarkable mountain range extending along all of the western side of

the island, with a structural alignment NE-SW, mostly characterized by rough landscapes and

well-karstified carbonate bedrock. The highest point on the island, Puig Major (1,445 m), is

localized in the middle section of the Tramuntana massif. The territory has been modified by

human activities over time as testified by the presence of terraces and farmlands spread all over

the region (Ginés, 1999). Serres de Llevant is the second important orogenic range of

Mallorca, lengthened in the eastern region along a NE-SW direction parallel to the coast. The

ridge consists predominantly of rounded and gentle promontories, shaped by complex thrust

5

structures (Gelabert, 1998). Es Pla, is a depression extending between the two mountain belts

and generally characterized by a flat landscape. In this area consisting of recent colluvial

deposits and red soils, is where most of the agriculture activities of the island are located.

Apart from these principal structural units, other areas have relevance for their geomorphologic

characteristics. The central part of the depression, for instance, is marked by a hilly landscape

known as Serres Centrals, whose outcropping is connected to post-orogenic extensional

features. Furthermore, in the southeastern regions of Migjorn and Llevant, carbonate platform

deposits crop out along the coastline in outstanding cliffs interrupted by morphologic incisions

locally known as “calas” (bays) (Gómez-Pujol et al., 2013).

Figure 1. Localization of Mallorca Island in the western Mediterranean. It is visible the structural

alignment NE-SW of the Balearic Promontory and the surrounding basins.

Image modified from Google Earth.

6

2.1.2 Climate

The island of Mallorca is characterized by a typical Mediterranean climate showing two

distinct seasonal trends. Dry and hot summers alternate with mild-humid winters, with an annual

temperature mean of about 17°C. The average annual precipitation reaches 570 mm/yr, with

most of the rainfalls recorded in the Tramuntana range and less in the interior, eastern, and south-

eastern coast of the island (Guijarro, 1995). The rainfall events show an irregular pattern

throughout the year, being concentrated particularly during the winter period (Sumner et al.,

2001). In the past decades, increasing episodes of extreme precipitation events and storm surges

have been recorded at local and regional scale, predominantly during late summer and autumn.

Occasionally, the magnitude of these events were rather significant causing serious infrastructure

damages especially in the coastal areas, where most of the economic activities on the island are

concentrated (Ramis et al., 2009).

2.1.3 Geologic Settings

The geologic structure of Mallorca has been heavily affected by complex tectonic events,

principally thrusts and folds, connected to the Alpine orogenesis (Fornós and Gelabert, 1995).

The stratigraphic succession includes rocks formed from the Triassic to the Quaternary period,

with a hiatus between Paleocene and lower Eocene (Fig. 2). The Mesozoic section (Triassic-

Upper Cretaceous) is well exposed in the thick cliffs of Serra de Tramuntana, as well as in the

promontories of Serres des Llevant, and the outcrops of Serres Centrals. The sequence consists

of micritic limestones with different textural characteristics that vary for each environment of

formation from shallow marine to pelagic, hemi-pelagic, lacustrine, and littoral (Rodríguez-Perea

and Gelabert, 1998). Gypsum and dolomites are also present in alternation with sandstones and

mudstones deposits, particularly in the Triassic and the early Jurassic periods. The Cenozoic

7

succession is predominantly composed of marine lithofacies, with little presence of continental

deposits (Fornós et al., 2002b).

Figure 2. Simplified geological map of Mallorca showing the principal morphostructural

and stratigraphic units (modified from Fornós et al., 2019).

The synorogenic sediment sequences (Eocene-Middle Miocene) are represented by

calcarenites and marls, interbedded with bioclastic limestones, which emerge principally in the

Serra de Tramuntana and Serres de Llevant regions and, sporadically in the central area of Es Pla.

The post-orogenic units (Upper Miocene) are made up of sequences of reefal limestones, bioclastic

calcarenites, and calcisiltites well exposed in the SE part of the island (Reef Complex Unit of

8

Llucmajor and Migjorn regions, Santanyì Limestone) (Fornós and Pomar, 1983a; Pomar et al.,

1996). The Plio-Quaternary deposits of calcisiltites and calcarenites are interlayered with strata of

carbonate conglomerates overlying most of the Es Pla Depression and the piedmont areas of Serra

Tramuntana (Inca-sa Pobla and Palma basins), as well as the southern part of the island (Campos

Basin) (Fig. 2). The Upper Miocene limestones of Llevant and Migjorn platforms are covered by

aeolian (eolianites) and colluvial deposits that accumulated during the quaternary glacial periods

when sea-level dropped down about 150 m (Clemmensen et al., 2001). These Pleistocene deposits

are primarilly composed of cross stratified bioclastic carbonate sand and terrigenous material,

interlayered with carbonate silt and matrix-supported breccias (Clemmensen et al., 1997).

Stratigraphic and geodynamic investigations documented slightly low post-orogenic tectonic

activity on the island, mainly consisting of Neogene – Quaternary extensional faulting and low

vertical movements (Giménez and Gelabert, 2002; Just et al., 2011). Thermal anomalies have been

also documented in the southeastern region of Llucmajor-Campos Basin (López, 2007).

2.1.4 Mallorcan Karst: Speleogenetic Processes and Cave Types

The carbonate lithologies of Mallorca have been affected by long-lasting karst processes

that generated distinctive dissolution morphologies, both at the surface (exokarst) and

underground (endokarst) as defined by Bögli (1980). The exokarst of Mallorca includes karren

landscapes, biokarst morphologies, canyons, and solutional depressions (Ginés and Ginés, 1995).

These are predominant in the karst areas of Serra de Tramuntana and Serres des Llevant,

although coastal karren and dissolution features such as pans (kaminitza) and grooves are also

present on the southeastern Miocene carbonate cliffs. On the other hand, the endokarst features

refers to conduits and caves generated in the subsurface carbonate environments by waters

flowing along various underground paths. In this context, the lithological and hydrogeological

9

settings, as well as the climatic conditions, control most of the speleogenetic

processes/mechanisms. On the island of Mallorca, Ginés and Ginés (2011) described five major

categories of caves: vadose shafts, and vadose caves, phreatic caves, coastal caves, and

hypogenic caves (Fig. 3).

Vadose shafts are vertically-shaped conduits, particularly abundant at the high elevations

in the Serra de Tramuntana karst region (1a-c in Fig. 3). These are cavities formed by solutional

processes of descending meteoric waters in the vadose zone with preferential flowing along

joints and fractures in the karst massif (Ginés and Ginés, 1987).

Vadose caves refer to a series of air-filled, sub-horizontal cavities (mostly occurring in

the Tramuntana-Llevant karst ranges) consisting of galleries, extensive chambers, and canyon-

type passages (2a-b in Fig. 3). Their morphologic features are controlled by the

lithostratigraphic disposition of structural discontinuities such as bedding planes, joints, and

faults, as well as by the precipitation regime affecting the karst massifs (Palmer, 1991).

Breakdown processes are also very common in the vadose caves on Mallorca, with breakdown

blocks usually covered by outstanding speleothem decorations (Ginés and Ginés, 2011).

The phreatic-type caves consist of circular and oval-shaped passages formed below the

water table (saturation zone) (3a-b in Fig. 3). Network patterns, as well as horizontal conduits

with phreatic solutional morphologies like cupolas, are also very common features. Various

cavities developing both inland or near the coast have been included in this morphogenetic

category, particularly those located in the coastal phreatic areas of Tramuntana region.

Nevertheless, following variations in the hydrological settings, some of the phreatic caves have

undergone considerable morphological changes while exposed to vadose conditions (Ginés,

1995).

10

Coastal/Littoral caves represent the most distinctive endokarst features of many karst

areas of Mallorca (4a-b in Fig. 3), although they are particularly concentrated in the eastern and

southeastern region of Migjorn. In this part of the island, karstification took place synchronously

with the early diagenetic processes (eogenetic karst), due to the hydrologic interconnection

between the coastal (unconfined) aquifers and the marine groundwater, as well as the high

porosity of the Upper Miocene carbonate lithofacies (Ginés, 2000). In this context, the coastal

speleogenesis of Mallorca shows similarities with the Carbonate Island Karst Model described

by Mylroie and Mylroie (2007), but their flank margin cave model does not apply to this karst

(Ginés et al., 2013). Instead, Ginés et al. (2014) suggested the use of the Carbonate Coast Karst

Model to describe the complexity of the littoral endokarst processes developing on the island.

Figure 3. Idealized model of the different cave typologies developing in the island of Mallorca

(modified from Ginés and Ginés, 2011).

11

The chemical interactions between freshwater and saltwater in the coastal phreatic zone

(so-called mixing zone) generate undersaturated conditions with respect to the CaCO3, increasing

the dissolution process of limestone, causing the formation of initial solutional conduits (Ginés

and Ginés, 2007). Multiple sea-level variations over the past 6 million years led to a series of

conspicuous ceiling breakdown and speleothems precipitation, thus significantly influencing the

morphology of the littoral caves. During sea-level still stands, recurrent cave precipitates like

phreatic overgrowth on speleothems (POS) formed at the vadose/phreatic interface of brackish

sea-level associated cave pools, recording the past sea level positions through the Pliocene and

Pleistocene eustatic oscillations (Ginés et al., 1981; Dorale et al., 2010; Tuccimei et al., 2010;

Polyak et al., 2018; Dumitru et al., 2018a). Whenever the sea-level was below the caves, these

have been exposed to vadose conditions, favoring the precipitation of stalactites, stalagmites, and

flowstones. Later on, ceiling collapse episodes provided access into some of the caves and

caused major morphological changes (Fornós et al., 2019).

Recent investigations in some caves located in the southeastern part of the island, have

identified peculiar morphologies of hypogene origin (5 in Fig. 3), such as feeder-like vertical

conduits, rising solutional grooves, corrosion cupolas, and alveoli (Ginés et al., 2009). The

proximity of certain cavities to thermal active zones supports the possible role of ascending

waters from deep-seated aquifers through extensional faults in the speleogenetic processes

(Ginés et al., 2017).

2.1.5 Paleokarst Features: Structural and Morphogenetic Aspects

In the eastern littoral cliffs of Mallorca abundant paleokarst structures can be observed.

The cliffs are built up of carbonate deposits accumulated in different depositional environments

during the Upper Miocene and consist of three principal lithologic units: the Calcisiltites Unit

12

with Heterostegina (carbonate platform), Reef Complex Unit (coral reef barriers and lagoons

with patch reefs) and Santanyí Limestone (shallow marine platform with mangrove marshes,

sand plains, and stromatolites) (Esteban, 1979; Fornós and Pomar, 1983b, 1984; Pomar, 1991).

Although, the paleokarst features are mostly located in the Santanyí Limestone and the upper

layers of the Reef Complex units, and according to Fornós (1999) and Robledo and Pomar

(2000), the genesis of the karst phenomena started with the initial dissolution of the aragonitic

minerals constituting the coral patches (Reef Complex Unit) in the mixing zone. Later on,

during sea-level low stands, aerial exposure of the cavities may have contributed to further

karstification processes.

Successively, sea-level rise and high stands throughout the late Miocene allowed the

deposition of shallow carbonates (Santanyí Limestone) on top of the reef barrier deposits. This

may have triggered the initial subsidence of the uppermost strata and partial collapses that

affected parts of different cavities. Finally, the total breakdown of the cave ceilings took place

with compaction and brecciation of the sediments filling the voids (Fornós et al., 1995) (Fig.

4A).

The degree of cementation and fracturing of the strata involved in the collapse process

may strongly affect the structure and geometry of the paleokarst features, as well as the type of

karst deposits. Within this context, two different categories of paleokarst structures have been

described: syn- and post-sedimentary (Robledo et al., 2004). In the former, the deposits were not

well cemented during the process defining the ductile deformation of the uppermost layers

(Santanyí Limestone) with funnel geometries. The cavities are filled mostly with non-matrix-

supported oligomictic breccia and include centimetric to metric sub-angular blocks (Fig. 4B-C).

13

Figure 4. A) Schematic summary of the paleokarst development in the Upper Miocene limestones

(modified from Fornós, 1999). Syn-sedimentary paleokarst structures: B) The eastern coastal zone of

Portocolom village, and C) Punta des Savinar (Mondragó). Post-sedimentary paleokarst features:

D) Littoral area near the entrance of Cova del Dimoni (Porto Cristo), and E) Cova des Pont (Manacor)

(photo credit: B.P. Onac).

14

In the post-sedimentary paleokarst, the upper strata were consolidated and fractured while the

collapse took place with blocks displacement involving the uppermost Santanyí Limestone, as well

as the Reef Complex lithologies. Crackle-chaotic-matrix supported breccia and angular boulders of

metric dimensions, filled the caves (Fig. 4D-E). Both types of paleokarst structures have undergone

morphologic evolution due to coastal dynamics and subaerial weathering (Balaguer, 2005).

The morpho-structural processes consist of different phases: 1) differential erosion of the

lateral parts of the collapse deposits (breccia) and mechanical displacement of the lateral blocks (a

in Fig. 5); 2) formation of littoral arches, natural bridges, and caves by erosion/corrosion processes

in the mixing zone (coastal caves) (b and c in Fig. 5); 3) breakdown of the karst features (d in Fig.

5), and 4) debris washout along the coastline and platform retreatment (e in Fig. 5) (Rosselló et al.,

2007). Intraparticle cementation may have occurred either in the phreatic or vadose conditions.

Figure 5. Morphological evolution of paleokarst structures in the eastern coastal region of Mallorca

(Solimina, Mondragó Natural Park) (photo credit: J. Ginés) (see text for explanations).

15

In some locations of the eastern coastal area, the post-sedimentary paleokarst structures

show distinctive features that are related to the endokarst processes in the vadose zone. The

speleothems exposed at these sites, especially stalagmites and flowstones, have been

precipitated when the paleo-coastal caves were air-filled (free of water) and meteoric processes

had a major contribution in the endokarst development. Most of these exposed speleothems

appear particularly old, thus potentially providing key information regarding the antiquity of

the karst and caves on Mallorca. To date, no geochronological studies of speleothems

preserved in paleokarst structures have been conducted on the island.

2.2 U-Th-Pb Geochronology of Speleothems

2.2.1 U, Th, and Pb in Natural Environments

Uranium is an element belonging to the actinide series of the periodic table with atomic

number 92 and two prevalent oxidation states, U4+ and U6+. The following three main isotopes

238U, 235U, and 234U, with abundances of 99.27%, 0.72%, and 0.0055% respectively, are present

in nature. Redox potential and pH variations, along with temperature and pressure mainly affect

uranium mobility in crustal fluids, as well as precipitation of uranium-bearing compounds in

sedimentary environments (Shock et al., 1997). In near-surface ground waters, U6+ is more

prevalent and usually bonded with two oxygen atoms to form the polyatomic uranyl cation

(UO22+), which may occur in hydrate complexes and as carbonate species under more oxidizing

conditions (Langmuir, 1978). In contrast, U4+ is extremely insoluble in waters with pH ranges of

6-8.5, although present in hydroxyl complexes and hydrated phosphates in highly reducing

environments. In carbonate minerals, Extended X-Ray Absorption Fine structure (EXAFS)

analyses demonstrated UO22+ substitution of Ca2+ and axial oxygen replacement of the (CO3)

2-

compounds in the calcite lattice, producing little structural deformation (Kelly et al., 2003). U4+

16

replacing Ca2+ have been also described in different calcite samples as being particularly stable

in the crystal structure, a reasonable condition due to the similarities of the ionic radius (Sturchio

et al., 1998; Cole et al., 2004).

Thorium is an actinide element of atomic number 90, with only one oxidation state, Th4+

and two naturally occurring isotopes, 232Th (~99.98 %) and 230Th (~0.02%). In the Earth’s crust,

thorium is found as a trace element in solid compounds like phosphates, silicates, and oxides, as

well as occluded in soil colloids and clay minerals (Langmuir and Herman, 1980). Although

mostly insoluble in surface and groundwater under the oxide form ThO2, Th mobility increases

when complexed with sulfates, carbonates, hydroxides, and organic molecules. Its presence in

the crystal lattice of carbonate minerals is related to detrital sources (terrigenous), and occurring

as a radiogenic product.

Lead is a metallic element (Group IVA) of atomic number 82 having two oxidation

states, Pb2+ and Pb4+, and four natural stable isotopes, 204Pb, 206Pb, 207Pb, and 208Pb, with

abundances of 1.4%, 24.1%, 22.1%, and 52.4%, respectively. In the natural systems, most of the

lead has oxidation state Pb2+. Dissolution of Pb in normal conditions of temperature and

pressure does not occur, although dissolved forms may be associated with carbonates [PbCO3,

Pb(CO3)22-] in acidic water solutions. Despite the low concentration of Pb in calcite minerals

(ppb), ionic substitution of Ca2+ in the octahedral coordination has been observed, showing low

distortion of the crystal lattice and relatively high stability (Reeder et al., 1999).

2.2.2 Overview of Radioactive Decay Mechanisms and Laws

Radioactivity is the spontaneous process of transformation of an element into another by

the emission of energy. Based on the type of subatomic particles emitted by the element

undergoing radioactive decay, three principal forms can be described: alpha (α), beta (β) and

17

gamma (γ) decays; nevertheless, other modes also occur such as electron capture, spontaneous

fission, and neutron emission (White, 2015). In α-decay, unstable nuclides (parent) disintegrate

to a radiogenic nucleus (daughter) releasing two protons and two neutrons (α particle), thus the

mass and atomic numbers decrease for the parent (A-4 and Z-2) in this transformation. In β-

decay, a neutron of a parent nuclide is converted into a proton with the release of an electron (β-

particle) plus an electron antineutrino without changing of mass number in the radiogenic

daughter. Likewise, a proton can convert into a neutron emitting a positively charged electron

(β+ particle) and an electron neutrino. In γ-decay, a nuclide energetically excited, resulting from

previous decay or a reaction, releases high energetic electromagnetic radiation (gamma rays)

transforming the nuclide into a lower, and more stable, energy state.

The number of unstable nuclides (N) that disintegrate per unit of time (t) is directly

proportional to the amount of the nuclides present at any stated time (universal law of

radioactivity) (Rutherford and Soddy, 1902):

𝑑𝑁

𝑑𝑡= 𝜆𝑁

(1)

where λ is the decay constant, which defines the probability that any radionuclide atom will

decay within a given time.

By integrating the equation (1) and rearranging the values, a new equation is obtained:

𝑁 = 𝑁0𝑒−𝜆𝑡 (2)

The rate law is related to the half-life (t1/2) of a specific parent isotope, which is defined as the

time necessary for half of the radionuclide to decay into a daughter isotope, thus depending on

the decay constant according to the equation below:

𝑡1/2 =

𝑙𝑛2

𝜆=

0.693

𝜆

(3)

18

In a closed system, the radiogenic daughter number (D*) forming at any time t is equal to the

number of parent isotope expired (N0-N). At t = 0, the daughter isotope value is defined as D0.

Therefore, the overall number of radiogenic nuclides (D*) at a given time t is calculated as:

𝐷 = 𝐷0+𝑁(𝑒−𝜆𝑡 − 1) (4)

which is the cornerstone equation (age equation) of most of the radiometric dating techniques

(U-series, U-Pb, Rb-Sr, etc.). Details for all the equations, steps, and geochronological

applications can be found in Dickin (2005).

2.2.3 U-Th-Pb Decay Systems and Dating Methods

The isotopes of uranium and thorium, 238U, 235U, and 232Th, are moderately unstable in

the Earth’s crust and undergo complex, long-lasting decay chains to form the stable radiogenic

isotopes of lead (206Pb, 207Pb, and 208Pb, respectively), known as uranium, actinium, and

thorium radioactive series (Fig. 6).

Figure 6. Uranium (green), Actinium (yellow), and Thorium (orange) decay series. Intermediate

radioactive daughter and half-lives (>10 years) are shown. Decay constants of 238U, 235U, and 232Th

are dispayed (from Schoene, 2014).

19

238U has a half-life of 4.4683×109 years and decay constant of 1.55125e-10 yr-1,

whereas 235U half-life is 7.0381×108 years and decay constant of 9.8485e-10 yr-1 (Jaffey et al.,

1971; Schoene et al., 2006). 232Th half-life is 1.405 ×1010 years, having a constant decay of

4.475 e-11 yr-1 (Steiger and Jäger, 1977). The less abundant isotopes 234U and 230Th are

intermediate decay products of the parent 238U, with half-lives of 2.4562×105 and 7.5584×10-4

years, and decay constants of 2.8221 e-6 yr-1 and 9.1705e-6 yr-1, respectively (Cheng et al.,

2013). The geochronology community regarding the decay constant values and relative

uncertainties continues to search for further improved values (Boehnke and Harrison, 2014;

Villa et al., 2016).

In U-Pb dating, ages can be calculated using more than one independent radioactive

decay system, and the three most important ones are 238U-206Pb, 235U-207Pb, and 232Th-208Pb

(Schoene, 2014):

𝑃206 𝑏

𝑃204 𝑏= (

𝑃206 𝑏

𝑃204 𝑏)

𝑖

+ (𝑈238

𝑃204 𝑏) (𝑒−𝜆238𝑡 − 1)

(5)

𝑃207 𝑏

𝑃204 𝑏= (

𝑃207 𝑏

𝑃204 𝑏)

𝑖

+ (𝑈235

𝑃204 𝑏) (𝑒−𝜆235𝑡 − 1)

(6)

𝑃208 𝑏

𝑃204 𝑏= (

𝑃208 𝑏

𝑃204 𝑏)

𝑖

+ (𝑇ℎ232

𝑃204 𝑏) (𝑒−𝜆232𝑡 − 1)

(7)

where 204Pb is the only non-radiogenic isotope of lead and is commonly used as reference in

the age calculation. In considering the reliability of the U-Pb radiometric dating, some

assumptions need to be made: 1) the system (mineral) must have remained closed to both

parent (U) and daughter (Pb) isotopes, as well as to all the intermediate radionuclides since the

time of crystallization and throughout successive perturbations (U loss, secular disequilibrium,

Pb loss/gain); 2) if present, initial Pb (common Pb) must be considered in the age calculation;

20

3) precise and accurate values of the 238U and 235U decay constants must be known, and 4) the

parent and daughter isotopic concentrations must be accurately measured with no systematic

errors (Faure and Mensing, 2005).

However, natural processes such as erosion, diagenesis and/or melting may cause U-

system perturbation and some intermediate decay products can be added or removed to/from

the system generating a condition where the decay activities of the parent and daughter

radioisotopes are not in equilibrium (activity ratios different from 1). When these fractionation

events take place, the system is reset to different disequilibrium states, known as daughter

excess (activity ratios > 1) and daughter deficiency (activity ratios < 1). In the former, a

daughter radioisotope (i.e. 234U) is initially present in large amounts with respect to its parent

(i.e., 238U). Through time, the nuclide excess is reduced by radioactive decay until the secular

equilibrium is reached, and the extent by which this process occurred, can be used as a dating

tool. In the daughter deficiency system, the parent radionuclide is independently deposited

with no radiogenic daughter present. With time, the amount of daughter increases until the

secular equilibrium is established. By measuring how much of the radiogenic product is

formed, the age of the mineral can be calculated (Dorale and Richards, 2003). This method,

known as 238U-234U-230Th disequilibrium or U-series dating, is one of the widest techniques

applied in speleothems’ geochronology (Dorale et al., 2004).

2.2.4 U-Pb Age Determination Methods

A widely used approach for U-Pb age calculation is the two-dimensional (2D) isochron

method, which plot the ratio 238U/204Pb or 235U/204Pb on the x-axis and 206Pb/204Pb or

207Pb/204Pb on the y-axis. This procedure allows to determine the initial ratio of common Pb

(Pbi) and to define the age of the sample analysed by using linear regression (the slope of the

21

line is time dependent). Three dimensional (3D) isochrons are also used for U-Pb system to

precisely define the age of samples formed at the same time (co-genetic), as well as the initial

ratio of 206Pb/204Pb and 207Pb/204Pb (Zheng, 1992). One of the advantages of using the 3D

isochron method in which 238U/206Pb vs. 207Pb/206Pb and 204Pb/206Pb (third axis) are

simultaneously plotted, is that it allows the estimation of Pb gain or loss in/from the system

(Ludwig, 1998). Because of differences between the decay system of 238U and 235U, cross-

checking of age suitability can be defined using the Wetherill Concordia (Wetherill, 1956) and

the Tera-Wasserburg Concordia methods (Tera and Wasserburg, 1972), plotting the 207Pb/235U

(x-axis) vs. 206Pb/238U (y-axis), and 238U/206Pb vs. 207Pb/206Pb, respectively. These methods are

useful to detect the period during which the samples remained in a closed system, i.e., since the

calcite (or aragonite) precipitation took place (concordant ages), as well as the time when

system perturbation occurred (discordant ages). Concordant ages lie on the concordia curve,

whereas the discordant ones fall outside the curve; a discordia line intercepting the curve

determines the time of mineral formation and system alteration. Furthermore, in the Tera-

Wasserburg Concordia, the x-y values error correlation is minimized, and do not require

common Pb (Pbc) isotopic composition for the age calculation, which can be determined by

combining 3D isochrons to the Concordia method (Wendt, 1984; Ludwig, 1998).

2.2.5 Previous U-Pb Dating Studies of Speleothems

Secondary cave deposits, particularly flowstones and stalagmites, have been widely

used in the past decades as exceptional archives for paleoclimate studies, recording

information about past variations in temperature and humidity, tectonic processes, vegetation,

sea level, as well as landscape and human evolution (Gascoyne, 1992; Richards and Dorale,

2003; Fairchild and Baker, 2012). Furthermore, radiometric dating techniques, especially U-

22

series, demonstrated the extraordinary suitability of speleothems for geochronological

investigations, although showing age accuracy limitations in samples older than ~500 kyr

(Ludwig et al., 1992; Dorale et al., 2004). In this context, U-Pb dating method, first applied on

carbonate deposits by Moorbath et al. (1987), has been extended to the speleothem realm

allowing researchers to document karst-based paleoclimate and geomorphology records

beyond the U-Th boundaries (Richards et al., 1998; Lundberg et al., 2000; Woodhead et al.,

2006; Polyak et al., 2008).

In their pioneering work, Richards et al. (1998) analyzed a stalactite from Winnats

Head Cave (UK), using thermal ionization mass spectrometry (TIMS) for U/Pb ratio

measurements and isotope dilution (228Th/232U spike) for U and Th determination. Although

the sample revealed a relatively young age (248 ± 10 ka), the investigation proved the

successful application of U-Pb dating technique in the study of cave formation. Furthermore,

Richards et al. (1998) denoted that for speleothems younger than 1 Myr, accurate correction

for initial isotopic disequilibrium (intermediate radiogenic isotopes) needs to be made.

Soon after, Lundberg et al. (2000) performed U-Pb analyses on a calcite spar sample

from Big Canyon in the Guadalupe Mountains National Park (USA). Isotopic ratio

measurements were carried out by using TIMS, whereas the age calculations were obtained

using the isochron method and Concordia plot. An estimated calcite spar age between 87 and

98 Myr suggests the occurrence of hydrothermal-driven karst process during the Upper

Cretaceous, along with the uplift of Guadalupe Mountains during the Laramide Orogeny.

Woodhead et al. (2006) applied the U-Pb method by using multi-collector inductively

coupled plasma mass spectrometry (MC-ICPMS) technique to measure the U/Pb ratios in

different speleothems from Leana’s Breath Cave (Nullarbor, Australia) and Antro del Corchia

23

(Alpi Apuane, Italy). In this study, the analytical issues denoted by Richards et al. (1998),

such as mass fractionation and impurity sensitivity, were partially solved by the

instrumentation resolution improvement of the MC-ICPMS, coupled with the use of a mixed

tracer 233U-205Pb.

U-Pb dating performed on flowstone samples (interbedded with hominid fossil-bearing

deposits) from different cave sites in the “Cradle of Humanity” region in South Africa

(Sterkfontein, Swartkrans, Coopers and Malapa), revealed new insights in the Pre-Olduvai

human history (Walker et al., 2006; Pickering et al., 2019). Uranium presence in the

flowstones’ layers were pre-examined using the beta scanning method (digital

autoradiography), useful to improve the preciseness of the technique by selecting calcite within

the stratigraphy having more abundant concentrations of uranium (Pickering et al., 2010).

Polyak et al. (2008) investigated speleothems that characteristically form near the water

table (mammillary) located in some caves in the Grand Canyon National Park (USA), using U-

Pb dating analysis. These usually precipitate along the cave walls and roofs in phreatic

environments due to high saturation conditions of water with respect to the CaCO3 (Hill and

Forti, 1997). The U-Pb dates of the cave deposits yielded the timing and position of the water

table in different locations within the Grand Canyon. This permitted the interpretation of

Grand Canyon incision history and defined different incision rates for the western (55 to 223

m/Myr) and eastern (166 to 411 m/Myr) sectors, respectively.

Besides these high-profile investigations, peer-reviewed studies using U-Pb dating of

speleothems have been published in the last few decades, confirming the potential

opportunities (and challenges) of the method in different geological research fields (Cliff et al.,

2010; Polyak et al., 2016; Decker et al., 2017). Detailed reviews of the U-Pb geochronology

24

strategies of application to old cave formations, as well as the methodological and analytical

aspects can be found in Woodhead et al. (2012) and Woodhead and Pickering (2012).

2.2.6 U-series Age Determination

U-series dating method of secondary cave precipitates relies on the basis that uranium

radionuclides (238U and 234U) are highly fractionated (chemically and/or physically) from the

radiogenic daughter 230Th in the hydrosphere, due to the high solubility of the firsts compared

to the second (Dorale and Richards, 2003). However, in water-bedrock interactions 234U is

preferentially carried in the fluid phase than 238U, resulting in a 234U-238U disequilibrium at the

moment of cave calcite (or aragonite) precipitation. By considering these initial disequilibrium

conditions, U-series age of the speleothems can be calculated using the following equation

from Kaufman and Broecker (1965):

[

𝑇ℎ230

𝑈238 ] = 1 − 𝑒−𝜆230𝑇 + (𝛿234𝑈(𝑚)

1000) (

𝜆230

𝜆230−𝜆234) (1 − 𝑒(𝜆234−𝜆230)𝑇)

(8)

where λ230 and λ234 represent respectively the decay constants of 230Th and 234U, [230Th/238U] is

the activity ratio of 230Th/238U, and δ234U(m) is determined from the activity ratio [234U/238U] by

the relationship: ([234U/238U] - 1) × 1000 (Edwards et al., 1987). The activity ratios 230Th/238U

and 234U/238U are measured using different mass spectrometric instrumentations (TIMS, MC-

ICP-MS, etc.), and when these values are known, the equation can be solved for time (T).

Even though, the equation assumes an ideal initial absence of 230Th, some non-radiogenic 230Th

can be trapped in the cave calcite (or aragonite) minerals and initial 230Th correction is

necessary in the age calculation (Hellstrom, 2006).

25

2.2.7 Previous U-series Investigations of Secondary Cave Deposits

Although the discovery of uranium radioisotopes dates back to the end of the nineteenth

(Bequerel, 1896; Curie, 1898) and early twentieth century (Thomson, 1914), first attempts of U-

series geochronological studies were made only in the 1950’ and 1960’ (Barnes et al., 1956;

Rosholt and Antal, 1962; Cherdyntsev et al., 1965). Only in the 1970’ the U-series techniques

began to be applied as a routine technique to date secondary cave deposits (Ford et al., 1972;

Thompson et al., 1974, 1975; Harmon et al; 1975). However, major issues about the age

accuracy were concerning the ability of measure such low isotopic concentrations in the samples.

Analytical improvements were gained in the late 1980’ with the advance of thermal ionization

mass spectrometry (TIMS) for 234U and 230Th measurements (Chen et al., 1986; Edwards et al.,

1987), and applied for the first time on speleothems by Li et al. (1989). Successively, the

development of the inductively coupled plasma mass spectrometry (ICP-MS) and multi-collector

inductively coupled mass spectrometry (MC-ICP-MS), allowed to precisely measure extremely

low isotopic concentrations (ppb-ppt), increasing the age accuracy of the method for samples

older than ~ 500 kyr (Luo et al, 1997; Cheng et al., 2000; Hellstrom, 2003; Andersen et al.,

2004). A detailed review of the analytical history, improvements and challenges can be found in

Cheng et al. (2013).

2.3 Oxygen and Carbon Isotope Geochemistry of Speleothems

Geochemical analyses of speleothems, and particularly the oxygen and carbon isotopic

compositions, have been widely applied in the past decades to investigate paleoclimate

conditions in terrestrial environments (Hendy and Wilson, 1968; Hendy, 1971; Schwarcz, 1986;

Gasconye, 1992; Dorale et al. (2002), McDermott (2004), Mickler et al., 2004; Fairchild and

Baker, 2012). Variations of oxygen (δ18O) and carbon (δ13C) isotopes in secondary cave calcite

26

reflect changes in the surface and groundwater chemistry, air-water interaction, air masses

circulation, as well as soil component and relative biological activity (Salomon and Mook, 1986;

Smart and Friedrich, 1987; Koster et al., 1993; Baskaran and Krishnamurthy, 1993). In

particular, changes of spelean δ18O values are significantly influenced by the so-called: i)

temperature effect, documented as a positive correlation between mean δ18O value and mean

annual temperature at the cave site; ii) amount effect that reflects an increased δ18O value of

rainwater with the decrease of rainfall amount; iii) ice volume effect, which is a consequence of

changes in isotopic composition of ocean water due to waxing/waning of ice sheets; iv) source

effect that causes δ18O values of rainfalls to vary when they have different moisture sources, and

v) the continental effect manifested by more negative δ18O values with increased distances from

the ocean (Dorale et al., 2002; McDermott, 2004; Mickler et al., 2004; Lachniet, 2009). δ13C

variations in speleothems are largely influenced by: abundance of C3 and C4 plants above the

cave, soil respiration and contribution of soil-derived and atmospheric CO2 to the dissolved

inorganic carbon (DIC), CO2 degassing in the epikarst (prior calcite precipitation; PCP) and

cave, and fluid-bedrock interaction (residence time) during water seepage from the surface to the

cave (Dorale et al., 1998; Mickler, 2004; Hartman and Baker, 2017). More detailed reviews on

oxygen and carbon isotopic variations in speleothems and their application as indicators of

environmental and climatic changes can be found in McDermott (2004); McDermott et al.

(2006); Fairchild et al. (2006), Lachniet (2009), and Fairchild and Baker (2012).

27

CHAPTER THREE:

MATERIALS AND METHODS

3.1 Investigation Sites and Samples Description

A total number of twenty speleothem samples, both flowstones and stalagmites, were

collected from different locations of the eastern region of Mallorca between 2014 and 2018

(Fig. 7). Most of the speleothems were exposed in paleokarst structures developed along the

coastline, at different elevations from the present sea level position. In addition to these, two

stalagmites samples (CF-11 and DR-D20) were recovered respectively from Cova de Cala Falcó and

the worldwide renown touristic attraction Coves del Drac. All the sites investigated in this research

project are shown in Figure 7.

Figure 7. Map of Mallorca showing the location of the sampling sites: Cala Falcó (CF), Coves del Drac

(DR), Dimoni (DIP), sa Ferradura (FE), and na Mitjana (MIP). All locations are color coded.

28

Cala Falcó site (hereafter CF) is a small bay located in the Manacor area, at about 2 km

south of the touristic location of Estany den Mas (Cala Romantica). The cala, like most of the

eastern rocky coastline inlets of Mallorca, developed in the Upper Miocene limestone from the

combined action of coastal erosion and hydrographic incision (Gómez-Pujol et al., 2013). A

conspicuous quantity of meter-size boulders overspread the shoreline, in proximity of which, the

homonym cave (Cova de Cala Falcó) is situated (Fig. 8A). Near the current cave entrance,

collapsed blocks cropping out on the cliffs are covered by a series of cave formations, in which

well-preserved flowstones and stalagmites can be observed (Fig. 8B, E). Sampling campaign

took place between 2017 and 2018.

Figure 8. Cala Falcó site: A) Overview of the cala and the cave entrance; B) Panorama of the paleokarst

structure outcropping on the cliff; C-E) Close up of the sampling points

(photo credits: B.P. Onac and J. McIlrath).

Samples CF-10 and CF-11 were collected a few meters from the cave entrance: CF-10

represents the lowermost part of a well layered, crystalline flowstone unit exposed on the

29

roof of the cavity at ca. 1.5-2 m above present sea level (mapsl) (Fig. 8C); the 4.5 cm tall

stalagmite CF-11 was recovered at ~ 1 mapsl from the upper section of a flowstone formed on

top of a red matrix carbonate sand-size deposit in which fragments of Myotragus sp. bones were

observed (Fig. 8D). The stalagmite was partially broken and highly weathered on its upper

surface. Two samples collected in the paleokarst structure exposed on the cliff outside the cave

from a ~ 23 cm thick flowstone; samples CF-12 and CF-13 represent the inner and lower

portions of the flowstone unit, respectively, whereas CF-14 is the uppermost part of the entire

speleothem section (Fig. 8E). All samples appeared nicely bedded with layers of translucent

compact calcite that alternate with ‘dirty’ calcite strata with detrital presence.

Dimoni site (DIP) is an embayed coastline landform located at ca. 1 km NE of the Porto

Cristo municipality, characterized by polymictic breccia deposits cemented in a red matrix of

carbonate sand and silt (Fornós, 1999) (Fig. 9A-B). Within these sediments, a series of

secondary cave formations of vadose origin, mostly flowstones and stalagmites, are present.

These cave precipitates, along with the breccia deposits, testify to the presence of coastal paleo-

caves, which may have undergone breakdown processes following past sea level oscillations.

Successively, coastal erosion affected the paleokarst structure and modeled the bay geometry.

Although several voids and minor cavities can be observed within these breccias, of great

interest, both scientific and touristic, is Cova del Dimoni. The cave develops in the Tortonian

(Upper Miocene) carbonate units and hosts extraordinary speleothem decorations and spectacular

cave pools directly connected to the sea via submerged passages.

Well preserved flowstones were identified and sampled during the 2014 sampling

campaign in two different locations of the paleokarst structure. DIP-1-A sample was

collected from a 20 cm thick flowstone located at about 5 mapsl. The flowstone is embedded

30

between centimeter- to meter-size cemented carbonate blocks and show well crystallized

columnar compact calcite layers alternating with red-brown, columnar open calcite rich in

detrital impurities (Fig. 9C). Sample DIP-1-B was recovered from an 8-cm thick flowstone

located at ca. 3 mapsl, which consists of cm-size layers of white crystalline calcite

interbedded with thin laminae of red detrital sediment (Fig. 9D).

Figure 9. A) Dimoni outcrop the breccia deposits with flowstones and stalagmites exposed, both

truncated and/or embedded; B) Close up of the red polymictic breccias; C) DIP-1-A and D: DIP-1-B

flowstones, respectively (photo credits: B.P. Onac and J. McIlrath).

Sa Ferradura site (FE) is located at about 500 m NE of Dimoni site. The area is

mostly composed of deposits of meter-to-centimeter size large blocks cemented in a red

carbonate sandy matrix. Embedded in the breccia sediments, a 2 m thick section of

31

flowstone outcrops between ca. 3.5 and 5.5 mapsl. The flowstone presents different

crystalline fabrics, from columnar open to compact, and/or fibrous, and its color varies from

white-grey to red-brown. Three samples were collected from this outcrop (sampling

campaign 2014): FE-1-A (10 cm, white, columnar open calcite) was recovered from the

lower portion at. 3.5 mapsl; FE-1-E (11 cm, dark brown, columnar compact calcite with

coarse crystals) comes from the middle part ~ 4 mapsl; and FE-1-F (7 cm, grey, compact-

microcrystalline calcite alternating with thin layers of detrital sediment) was sampled from

the top flowstone at ca. 5.5 mapsl (Fig.10).

Figure 10. The 2 m thick section of flowstone located in the sa Ferradura site. The sampling points are

shown: FE-1-A (lower), FE-1-E (middle), and FE-1-F (top) (photo credits: B.P. Onac).

Na Mitjana site (MIP) is situated in the northeastern coastline of Mallorca, at about 1

km N of the touristic Coves d’Artà. The bedrock geology of this area consists of Jurassic

32

limestones overlaid by Pleistocene eolianites (Fornós et al., 2012). Na Mitjana Cave

develops within the limestone unit and has its entrance located at ca. 6 mapsl. Surrounding

the cave, remnants of several paleokarst structures comprised of angular metric blocks

cemented in red carbonate sand matrix can be observed. Well preserved flowstones and

other secondary cave precipitates are embedded between boulders and/or in minor voids and

cavities (Fig. 11).

Figure 11. Na Mitjana site. A) Cemented blocks in which embedded speleothems are exposed; B) MIP-1

and MIP-2 sampling points; C) Small cavity on the right side of na Mitjana Cave entrance; D) Close up of

MIP-3 sampling point; E) Close up of the flowstone from which MIP-4-A and MIP-4-B were collected;

F) View of the MIP-5 sample (photo credits: B.P. Onac).

During the 2014 sampling campaign, a total of five samples were collected from different

points of the paleokarst outcrops. Samples MIP-1 and MIP-2 were recovered from a 27 cm thick

flowstone at ca. 7 mapsl. MIP-1 was part of the lower flowstone portion and appeared nicely layered

33

with compact calcite crystals with colors ranging from white to dark-brown, whereas MIP-2 was a 4

cm tall stalagmite formed on top of the flowstone section, 35 cm away from MIP-1 (Fig. 11B).

Samples MIP-3, MIP-4A, and 4B were collected from a cavity located at ca.10-11 mapsl,

on the right side of the Na Mitjana Cave entry (Fig. 11C). MIP-3, consisting of fascicular-fibrous,

white-gray calcite crystals of 3 cm in thickness, was sampled from a mammillary-like speleothem

coating the walls and roof of the void (Fig. 11D). The base of the cavity, at ca. 10.2 mapsl, hosts a

21 cm thick dark-brown flowstone composed of columnar compact calcite, in which two

discontinuities were observed at 5.5 and 11 cm in depth. Samples MIP-4-A and MIP-4-B were

collected from the base and top of the flowstone unit, respectively (Fig. 11E). A few meters above

the Na Mitjana Cave entrance another cavity is present (~ 12 mapsl); from this one, sample MIP-5

was recovered (Fig. 11F). The speleothem growing from the roof of the void showed similarities

to the POS already described in other caves of the island by Ginés et al. (1981), Fornós et al.

(2002a), and Ginés et al. (2012). Two main crystallization patterns were observed on the 7 cm

thick MIP-5 sample: an inner part composed of translucent columnar-crystalline calcite layers

interrupted by thin white inter-strata; and an outer section characterized by brownish, columnar

open-compact calcite laminae interbedded with detrital-rich thin red layers (Fig. 11F).

Lastly, a 10 cm tall stalagmite was collected from Llac Delícìes located in the

easternmost area of the touristic Coves del Drac, very close to the present-day seashore (Fig.

12A and B). It was recovered from ~ 1.1 m below present sea level and appeared highly

weathered on its outer surface (Fig. 12C). However, the cross-section of the sample revealed

distinct morphological features (Fig. 12D). A soda straw was identified in the central part,

surrounded by translucent calcite layers parallel to the growth axis (stalactite) (Fig. 12D). The

smaller part (stalagmite) shows translucent calcite crystals of dark grey color, precipitated on

34

the side and partially englobing the stalactite component. On the upper section, thin white

nicely banded laminae of (probably) aragonite cover the speleothem.

Figure 12. A) Aerial picture of the Portocristo area and location of the Coves del Drac (from Ginés et al.,

2014), B) Image of the Llac Delícìes (photo credit: J. Ginés) , C) Stalagmite DR-D20 showing in its outer

surface weathered white layers; and D) Sample cross-section displaying soda straw, stalactite, and stalagmite

components with the thin white weathered laminae on the upper part.

The in-situ speleothems from Cala Falcó, Dimoni, Sa Ferradura, and Na Mitjana were

particularly well preserved, and indicate the presence of paleo-coastal caves showing evolution under

vadose conditions. The investigation of such cave deposits by means of U-series dating indirectly

imply lower sea level positions and constrain the timing when breakdown and dismantling of the

paleo-caves occurred.

35

3.2 U-series and U-Pb Dating Analyses

Dating analyses were performed in 2018 at the Radiogenic Isotope Laboratory, Department

of Earth and Planetary Sciences, University of New Mexico (Albuquerque, NM) under the

supervision of Drs. Y. Asmerom and V.J. Polyak.

3.2.1 Samples Preparation

All samples were carefully selected and cleaned before the chemical analyses. Wet

cutting procedures were performed using a trim saw with a RIDGID Diamond edge blade

(1.21 mm) and, where applicable, clean and crystalline calcite layers were extracted, cleaned,

and broken into smaller pieces. A few samples, such as CF-11, MIP-5, and DR-D20, were

sliced parallel to the growth axis and instead of pieces, powder sub-samples (0.04-0.12 g)

were drilled from different individual growth layers. Before drilling sample powders, all

surfaces were cleaned to remove external detrital particles from the mineral surface by using

compress air, washed with ionized water, and cleaned with methanol. Prior to column

chemistry, all sub-sample ‘pieces’ were etched for a few seconds with 3% HNO3 in a

Polytetrafluoroethylene (PTFE) cup, rinsed three times in ultra-pure deionized water (18

MΩ), and dried on a hotplate at a temperature of ca. 150-180°C (below PTFE melting

temperature).

3.2.2 Chemical Procedures

Dried etched pieces were wrapped in a clean napkin and crushed in a plastic bag using

a 3" Agate mortar and pestle (Fig. 13A). The smaller pieces were collected in 8 mm

(diameter) plastic tubes (Fig. 13B) and each piece selected for analyses was weighted using a

Mettler Toledo AT261 DeltaRange balance. To avoid any sample contamination throughout

the chemical processes, Teflon still distilled ultra-pure acids were used.

36

Figure 13. Sub-samples preparation and chemical proceedings. Sub-samples are crushed (A), and

stored in tubes (B); C) Dissolved aliquots are spiked, purified and dried down on hotplate;

D-E) U, Th, and Pb are chemically separated (see text for detailed explanation).

For U-series analyses, 0.05-0.15 g mineral aliquots were chosen and dissolved in 15N

HNO3. Solutions were spiked with a synthetic 229Th-233U-236U mixed tracer, fluxed for ~1-2

hours, and dried down at 180°C (Fig. 12C). Additional conditioning steps were made by

adding drops of 15N HNO3 and HClO4 to eliminate any presence of organic contaminants

and to equilibrate the spike with the sample.

For U-Pb dating, I used five mineral aliquots (0.05-0.2 g) for each sub-sample (plus one

blank). These were dissolved in concentrated HNO3, spiked with 229Th-233U-236U-205Pb

mixed solution, and fluxed for ~1-2 hours. Conditioning procedures included several steps

of drying at 180°C and dissolving in 6N HCl, 2N HCl, and 1N HBr.

U, Th, and Pb chemical separations were performed using Eichrom 1x8 chloride form

200-400 mesh anion resin and 2-ml anion exchange disposable polypropylene

37

chromatography columns (Fig. 13D, E). Cations (Ca2+, Mg2+, etc.) were flushed by making

precise loadings of 7N HNO3 for U-Th dating, and 1N HBr for U-Pb dating. U and Th were

collected with 1N HBr and 2N HCl, and Pb was collected with 6N HCl in the first column

set. Th and U were separated in a second set of anion resin columns and collected using 6N

HCl and H2O, respectively. Collected solutions were dried down and were analyzed in 0.4

ml of 3% HNO3.

3.2.3 Analytical Methods

U, Th, and Pb isotopes were measured in separate runs using a Thermo NEPTUNE

multi-collector inductively coupled plasma mass spectrometer (MC-ICPMS), which provided

a simultaneous analysis of up to nine distinct isotopes. Instrumental sensitivity has been

enhanced by a factor of 4 using a Cetac Aridus II desolvating nebulizer system. Ion signals

(counts per seconds [cps]) were measured using a secondary electron multiplier (SEM) and

signals larger than 0.5 E6 cps were measured in volts in the Faraday cups. The SEM detector

in the central position (C in Fig. 14A, B) was used to measure the isotopes of least abundant,

which were 234U and 230Th. The Faraday cups were set to measure the other more abundant

isotopes (L1-L4 and H1-H4 in Fig. 14A, B).

Three different resistors, 1010, 1011, and 1012 Ω, were linked to Faraday cups to

optimize the signal detection up to 500, 50, and 5 volts, respectively. For Pb measurements,

the central position (C in Fig. 14C) was assigned to a Faraday cup with 1012 Ω resistor,

whereas the other isotopic analyses (204Pb, 206Pb, 207Pb, and 208Pb) were assessed to Faraday

cup detectors coupled with 1011 Ω resistors (L1-L4 and H1-H4 in Fig. 14C).

38

Figure 14. Cup configurations and resistors settings on the Thermo NEPTUNE MC-ICPMS for the U

(A), Th (B), and Pb (C) isotopic measurements (modified from LaPointe, 2013).

3.2.4 Data Processing, Age Models and Calculations

The mass spectrometric analyses generated the isotopic ratios datasets; 233U, 234U,

235U/233U, 234U/235U, 235U/233U, and 238U/233U ratios were reported for uranium, whereas 229Th,

230Th, 230Th/229Th, 238U/233U, 230Th/232Th, and 232Th/ 229Th ratios were measured for thorium, and

204Pb, 205Pb, 206Pb/204Pb, 207Pb/204Pb, 208Pb/204Pb, 206Pb/208Pb, 206Pb/207Pb, and 206Pb/205Pb for lead.

For U-Pb age calculations, data reduction was performed using the PBDAT computer

program (Ludwig, 1993). Excess/deficiency of 206Pb and 207Pb from initial disequilibrium were

corrected from initial excess/lack of 234U/238U, 230Th/238U, and 231Pa/235U, respectively (see

Supplementary Material in Polyak et al., 2008). U-Pb ages were calculated following equations 5

and 6 from Schoene (2014).

H4 H4 238U 10

10 Ω H4

H3 238U 10

10 Ω H3 235

U 1011

Ω H3 208Pb 10

10 Ω

H2 236U 10

12 Ω H2 233

U 1012

Ω H2 207Pb 10

11 Ω

H1 235U 10

11 Ω H1 232

U 1011

Ω H1 206Pb 10

11 Ω

C 234U SEM C 230

Th SEM C 205Pb 10

12 Ω

L1 233U 10

12 Ω L1 229

Th 1012

Ω L1 204Pb 10

11Ω

L2 232Th 10

11 Ω L2 L2

L3 L3 L3

L4 229Th 10

12 Ω L4 L4

U configuration Th configuration Pb configuration

Thermo NEPTUNE Cup Settings

B CA

39

Isochron ages were determined using the computer program ISOPLOT (Ludwig, 2001),

and ages expressed as two dimensional (2D) and three dimensional (3D) isochron dates. Updated

decay constants for 235U and 238U are reported by Schoene et al. (2006).

U-series ages were calculated following the equation (8) from Kaufman and Broecker

(1965) and modified by Dorale et al. (2004). Updated 234U and 230Th decay constants are from

Cheng et al. (2013).

δ234U model age was generated for samples that lied outside the U-Th dating range (~500-

600 kyr) using the 234U/238U geo-chronometer. In creating the model, assumptions for initial δ234U

were needed and those estimations, taken from dateable samples, were used to calculate the ages

using the following equation from Edwards et al. (1987):

𝛿234

𝑈(𝑚) = 𝛿234

𝑈(𝑖)𝑒−𝜆234𝑇 (9)

Finally, a uranium-evolution curve was created using ISOPLOT computer program

(Ludwig, 2001), plotting the activity ratios of 234U/238U (y-axis) and 230Th/238U (x-axis), which was

useful to assess the uranium systematics of the samples. This allowed to check if the speleothems

analyzed showed evidence supporting that they have remained in a closed system since their

formation or have undergone any alteration process (erosion, diagenesis, etc.) through time. Age-

depth model for the stalagmite CF-11 was generated in the open source statistical software R using

the algorithm StalAge developed from Scholz and Hoffmann (2011).

3.3 Oxygen and Carbon Isotopic Analysis

A total of 72 powders (50-120 g each) were micro-milled at 0.5 mm distance along the

growth axis of stalagmite CF-11 using a SHERLINE (model 5410) vertical drilling device. The

sample was positioned on a mobile platform connected to a readout tachometer in order to ensure

a stable and precise/controlled spacing during the drilling. The calcite powders were

40

subsequently reacted with concentrated H3PO4 at 25C in airtight vials for 24 hours, and the CO2

generated from the reaction was measured for oxygen and carbon isotopic ratios. The analyses

were performed in the Stable Isotope Laboratory of the School of Geosciences (University of

South Florida) using a ThermoFisher Scientific Delta V Advantage isotope ratio mass

spectrometer (IRMS). The 13C/12C and 18O/16O values were reported as δ-values (δ13C and δ18O)

in ‰ (permil), and corrected to the Vienna Pee Dee Belemnite (VPDB) scale using repetitive

analyses of the NBS-19 and Old-Car standards. The replicate measurements showed

reproducibility of ± 0.06 for δ13C and ± 0.11 for δ18O.

41

CHAPTER FOUR:

RESULTS

This chapter presents the results of the U-series and U-Pb analyses performed on speleothems

from the eastern part of Mallorca. In the first section, samples are discussed individually on a

location basis, whereas in the second part, the ages and uranium systematics are collectively

presented. All the speleothems showed low uranium content, ranging between 30 and 400 ppb, and

low to moderately high detrital Th variations (from 20 to 31,600 ppt). The majority of investigated

samples revealed ages close to the U-series dating limit (~600 kyr) and exhibited high age

uncertainties. Apart from these, samples CF-11, MIP-5, and DR-D20 indicate younger ages,

between 200 and 80 kyr, and are presented in separate paragraphs.

4.1 Cala Falcó

The samples CF-10, CF-12, CF-13, and CF-14 (Fig. 15A-D) from Cala Falcó had uranium (238U)

concentrations, ranging from 256 to 388 parts per billion (ppb). Moderately high thorium (232Th)

contents (relative to 238U concentrations) of 1,259 and 6,934 parts per trillion (ppt) were measured in

flowstones CF-10 and CF-12, respectively, whereas Th concentrations vary between 20 and 30 ppt in

CF-13 and CF-14 flowstones (see Appendix A, Table A1). Soil and bedrock are the principal

sources of detrital Th in vadose speleothems, such as flowstones and stalagmites (Polyak and

Asmerom, 2009). Thorium (mostly 232Th, and subordinately detrital 230Th) is absorbed at the surface

of clays minerals, colloids, and/or organic matter into the soil-zone and transported by the water

through the bedrock fractures inside the cave, where it is trapped in the speleothem, along with 238U

and 234U at the time of calcite precipitation (Richards and Dorale, 2003). Higher 232Th

42

concentrations, as observed in the samples CF-10 and CF-12, may suggest a major contribution of

soil-derived detrital Th, whereas lower concentrations, like those measured in the samples CF-13 and

CF-14, may indicate a more bedrock-derived input.

Figure 15. Speleothem samples from Cala Falcó used for U-series dating. The yellow bands mark

the position and code of each subsample.

The U-series age of flowstone CF-10 was 465 (+26/-22) kyr. The CF-12 inner flowstone and

the uppermost part of the CF-14 outcrop in the paleokarst structure on the cliff outside the cave (Fig.

8B, E), yielded U-series ages of 558 (+75/-50) and 500 (+51/-38) kyr, respectively. For the CF-13

flowstone, two U-series dating attempts were made, but none yielded robust results. The 230Th/234U

activity ratios of CF-13 indicate an age older than ~650 kyr, which is the 230Th-234U dating limit.

Nevertheless, initial δ234U values of 111‰ and 117‰ obtained from our analyses allowed the

generation of a δ234U model age of 653 (+32/-30) kyr (following the Eq. 9 in Chapter 3) with the

assumption that the initial δ234U of all these older flowstones does not differ significantly.

43

4.1.1 Stalagmite CF-11

The 4.5 cm tall stalagmite CF-11 from Cova de Cala Falcó shows concentrations of U

between 225 to 393 ppb and high variation in Th contents from 72 to 10,177 ppt (see Appendix

A, Table A1). Five U-series ages, at 35, 21, 18, 13, and 10 mm from top-down (color-coded

bands; Fig. 16A) were measured; the lowermost subsample (CF-11-35) formed 85.7 ± 0.5 ka,

whereas the uppermost CF-11-10 is 84.4 ± 0.4 kyr. The intermediate laminae (CF-11-18 and

CF-11-13) revealed U-series ages of 85.5 (+1.19/-1.17) kyr and 84.6 ± 0.9 kyr, respectively.

Subsample CF-11-21 has an age of 84.1 (+1.7/-1.2) kyr, thus is out of stratigraphic order (black

band in Fig. 16A), but its error range makes it acceptable. The high-resolution stratigraphy of

the CF-11 stalagmite is shown in Figure 16B.

Figure 16. Stalagmite CF-11. A) Cross section of the sample showing its individual growth layers and

internal corrosion features. The subsamples (color bands) used for U-series dating are also marked. CF-

11-21 (black band) denotes the outlier age. The double white lines represent brief hiatuses (corrosion

layers) at 20 and 11 mm from the top, respectively. B) Age-depth model for CF-11 stalagmite

(see text for details).

44

Even though most of the 230Th in speleothems is originating from the 234U decay, small

amounts of 230Th (detrital) are initially incorporated in the calcite minerals at the time of precipitation

and must to be accounted for in the age calculation. The 230Th/232Th atomic ratio commonly used in

the literature for the initial 230Th-232Th age correction is 4.4 × 10-6 ppm, which is assumed from the

Bulk Earth 232Th/238U ratio at secular equilibrium of ~3.8 (Ivanovich and Harmon, 1993; Kaufmann,

1993). However, soil and bedrock sources may contribute different isotopic ratios to the initial

230Th/232Th in the samples, with the soil-zone having 230Th/232Th signature very small (~0.0000044

ppm) compared to the bedrock 230Th/232Th values (~100 ppm) (Dorale et al., 2004). Presuming that

‘detrital’ 230Th and 232Th are chemically equivalent, and 232Th can be used to monitor 230Th, the

initial 230Th/232Th ratios were calculated for all the subsamples following the equation shown in

Figure 17.

Figure 17. Plot of 232Th vs Initial 230Th/232Th correction for all the subsamples (color-coded) and their

corresponding absolute 2σ errors. The equation used to calculate the initial 230Th/232Th correction is also

shown. The x values were assumed for each subsample from different 230Th concentrations in accordance

with the stratigraphic order.

CF.11.13

CF.11.10

CF.11.35

CF.11.21CF.11.18

0.0E+00

2.0E-05

4.0E-05

6.0E-05

8.0E-05

1.0E-04

1.2E-04

1.4E-04

1.6E-04

0 2000 4000 6000 8000 10000 12000

Init

ial

23

0T

h/2

32T

h

232Th (ppt)

(𝑇ℎ230

𝑇ℎ232 )𝐼𝑛𝑖𝑡𝑖𝑎𝑙 = 0.000562 · 𝑇ℎ232 – 𝑥

45

As suggested by Polyak and Asmerom (2009), relatively young vadose speleothems with low

uranium content, exhibit very high sensitivity to this correction for initial 230Th concentrations. The

correction curve generated by plotting the calculated initial 230Th/232Th ratios vs 232Th, shows a non-

linear trend (Fig. 17). Initial 230Th/232Th variations showed an inverse correlation with the 232Th

content. Subsamples with high detrital 232Th (CF-11-21, CF-11-18, and CF-11-13) in this model

exhibited low 230Th/232Th values, indicating a major contribution of detrital Th from the soil zone,

whereas those with low 232Th concentrations (CF-11-35 and CF-11-10) exhibit higher initial

230Th/232Th, evidencing more detrital Th coming from the bedrock. Moreover, at 20 and 11 mm

from the upper part, two corroded (reddish) laminae were identified (white double lines in Fig. 16A),

suggesting potential hiatuses caused by corrosion/weathering processes at ca. 85 and 84.5 ka.

The δ18O values (blue curve in Fig. 18) range from -5.1 to -2.0 ‰ with an average of -3.2 ‰.

From 85.7 to 85.6 ka the oxygen isotope values show a sharp decrease from -3.07 to -5.01 ‰, then

increase to -2.5 ‰ before dropping back to -3.78 ‰ at 85.5 ka.

Figure 18. The 18 O and 13C time series for stalagmite CF-11.

46

Between 86.4 and 84.7 ka, the δ18O values fluctuate between -2.6 and -3.94 ‰ (mean of -

3.39 ‰), with negative peaks at -4.14 and -4.52 ‰ at 85.3 and 85 ka, respectively. Subsequently,

another rise of -1.50 ‰ is observed from 84.76 (-3.50 ‰) to 84.7 ka (-2.0 ‰). After this change,

δ18O values fluctuate between -3.52 and -2.35 ‰ from 84.6 to 83.7 ka, reaching -4.06 ‰ at 83.6 ka.

The carbon isotope values (orange curve in Fig. 18) range between -7.07 and -1.03 ‰ with a

mean of -4.87 ‰. The δ13C record exhibits an overall similar trend with the δ18O profile, although

showing larger fluctuations and higher values between the period 85-84.4 kyr and 84-83.6 kyr.

The highest δ13C values are recorded at 84.5 ka (-1.23 ‰) and 83.7 ka (-1.03 ‰), whereas the

lowest were measured at 85.4 ka (-6.86 ‰) and at 84.1 ka (-7.07 ‰). From 85.7 to 85 ka, the carbon

isotopic profile displays a gradual increase with small fluctuations. After the peak at 84.5ka, δ13C

values present a rapid decrease within a range of -4.74 ‰ between 84.4 and 84.1 ka, and suddenly

rise to -1.03 ‰ at 83.7 ka, before dropping abruptly to -5.23 ‰ at 83.6 ka.

4.2 Dimoni

Two subsamples (DIP-1A-1 and DIP-1A-2) from flowstone DIP-1A and one from DIP-1B

(DIP-1B-1) were analyzed by U-series dating (Fig. 19A, B); uranium concentrations vary between

313 and 458 ppb. DIP-1A subsamples exhibit moderate detrital 232Th content (1.1 ppb and 2.7 ppb),

whereas DIP-1B-1 shows very low detrital 232Th value (17 ppt) (see Appendix A, Table A1).

Although no robust U-series dates were obtained, the 230Th and 234U values indicate a

potential age beyond 600 kyr. However, the δ234U model ages for the samples followed the Eq. 9 in

Chapter 3. According to Ludwig et al. (1992) and Polyak et al. (2013), single speleothem samples

spatially distant may show a relative concordance of initial δ234U values over a long time. For the

DIP flowstones, initial δ234U values were assumed to be concordant with the nearby sample FE-1-F l

from sa Ferradura site, with δ234U variations of 45 ‰ (low) and 50 ‰ (high).

47

Figure 19. Dimoni flowstones: A) DIP-1A-1 and B) DIP-1B. Subsamples’ code and position

are also displayed (yellow marks).

The δ234U model ages for DIP-1A-1 and DIP-1A-2 were 646 (+56/-52) kyr and 603 (+62/-

57) kyr, respectively, and 613 (+81/-73) kyr for the subsample DIP-1B-1.

4.3 Sa Ferradura

The two attempts to U-Pb date the flowstone sample FE-1-A (subsamples FE-1-A-1and FE-

1-A-2; Fig. 20A), returned no consistent results. This is due to extremely low U concentrations (11

to 150 ppb), and comparatively high total lead (Pb) contents (10 to 540 ppb). Th values were

between 30 and 190 ppt (see Appendix A, Table A2). The high amount of common Pb (non-

radiogenic) corroborated with the low U content, prevented the 238U-206Pb and 235U/207Pb geo-

chronometers to generate consistent and reliable isochron ages. However, the average activity ratios

of 230Th/238U (0.90794 ± 0.0068) and 234U/238U (1.043 ± 0.0017) indicate values very close to, but not

in secular equilibrium, suggesting an age of less than 1 Myr for the sample. The inner and top

flowstones FE-1-E (Fig. 20B) and FE-1-F (Fig. 20C) exhibit U concentrations of 385 and 202 ppb,

and Th contents of 168 and 4,597 ppt, respectively (see Appendix A, Table A1). Sample FE-1-F

48

indicates a U-series age of 478 (+29/-25) kyr, whereas FE-1-E shows 234U/238U and 230Th/238U

activity ratios of 0.9991 ± 0.0014, suggesting secular equilibrium.

Figure 20. Flowstone samples from sa Ferradura site: A: FE-1-A; B: FE-1-E; and C: FE-1-F.

Subsamples location (in yellow) and related codes are also shown.

Therefore, no robust U-series age was determined for this sample. However, a δ234U model age

was attempted assuming a present δ234U value of 10 ± 1 ‰ and a variation of the initial δ234U

between 45 ‰ (low) and 60 ‰ (high). These values represent the measured and initial δ234U of

the younger sample FE-1-F. The obtained δ234U model age of 583 (+88/-85) kyr is in agreement

with the stratigraphy of the flowstone.

4.4 Na Mitjana

Samples MIP-1 (Fig. 20A) from na Mitjana show very low uranium concentrations of 80

ppb and thorium content around 650 ppt. The uranium value in MIP-2 (Fig. 21B) was about 139

ppb, whereas Th concentration was 3,040 ppt (see Appendix A, Table A1).

49

Figure 21. Na Mitjana samples: A) MIP-1; B) MIP-2; C) MIP-3; D) MIP-4A; and E) MIP-4B.

Subsamples position (in yellow) and code are indicated for each speleothem.

U-series analyses indicate an age of 451 (+46/-35) kyr for the flowstone MIP-1 and 515

(+74/-49) kyr for stalagmite MIP-2. The mammillary-like speleothem MIP-3 (Fig. 21C) had

very low U content (82 ppb) and somewhat elevated Th concentration (5,730 ppt). The amount

of uranium in the flowstone samples MIP-4A (Fig. 21D) and MIP-4B (Fig. 21E) was 176 and

115 ppb, and Th contents of 1,085 and 203 ppt, respectively (Appendix A, Table A1). MIP-4B

revealed a U-series age of 507 (+89/-55) kyr. For the samples MIP-3 and MIP-4A, δ234U model ages

were generated assuming initial δ234U values varying from 33 ‰ (low) to 37 ‰ (high). These

assumptions result from calculating the average initial δ234U values from MIP-1, MIP-2, and MIP-4B

(33 ‰) for the low end, and taking as higher δ234U limit (37 ‰) the initial δ234U of MIP-4B sample

was used. The δ234U model produced ages of 580 (+77/-69) kyr for the mammillary-like calcite

MIP-3, and 590 (+78/-71) kyr for the top flowstone MIP-4A.

50

4.4.1 Sample MIP-5

Five powder aliquots, corresponding to 3, 8, 16, 18, and 25 mm from the outer surface,

were milled-out from speleothem MIP-5 (Fig. 22). Their uranium concentrations vary between

35 to 271 ppb, whereas the Th content was relatively high (3.5 to 16.5 ppb; see Appendix A,

Table A1). The presence at 20 and 23 mm below surface of thin interlayers of white color,

suggest distinct intervals of calcite precipitation hiatuses. Four subsamples located above (3 to

18 mm) and one below (25 mm) the hiatuses were dated by means of U-series technique. Due

to the higher detrital 232Th contamination, age calculations were corrected for the initial

230Th/232Th using an atomic ratio of 2.0 × 10-5 ppm (higher than the bulk Earth ratio). MIP-5-

18 revealed an age of 296 ± 2.7 kyr and MIP-5-16 of 284 ± 2.8 kyr. The outer subsamples

MIP-5-8 and MIP-5-3 yielded U-series ages of 224 ± 1.5 kyr and 186 ± 1.5 kyr, respectively.

Figure 22. Cross-section of MIP-5 sample. Powder subsamples and codes are displayed in yellow.

Hiatuses at 20 and 23 mm (from the outer surface) are shown in white.

The innermost MIP-5-25 (drilled below the hiatuses) did not returned a finite age,

indicating that it is potentially older than 600 kyr. A δ234U model age of 1.08 ± 0.06 Myr was

generated for the subsample using assumed initial δ234U values of 249‰ and 303‰ from the

51

lowest and highest initial δ234U, respectively. These ages document a long period (1.08 to 0.3

Ma) during which calcite precipitation ceased.

4.5 Coves del Drac

4.5.1 Speleothem DR-D20

From a short, 8 cm sample (DR-20) collected for a Holocene sea-level study, a total of

11 U-series analyses were measured as follows: one powder from the center section of the soda

straw and 3 aliquots at 3.5, 12.5, and 16.5 mm away from the straw (Fig. 23). Only one

subsample (DR-D20-4S) comes from the outer part of the smaller stalagmite (Fig. 23 left).

Uranium concentrations varied between 135 and 371 ppb, whereas thorium content ranged

from 36 to 1,113 ppt (see Appendix A, Table A1). Age correction of detrital Th assumed an

initial 230Th/232Th atomic ratio of 4.4 × 10-5 ppm, which is higher than the widely used 4.4 ×

10-6 ppm and to be applicable to the large 232Th variation observed in the speleothem. These

four subsamples of DR-20 represent a stalactite with a soda straw inner core. The age of the

soda straw sample is 225 ± 1 kyr, and the stalactite subsamples produced ages of 221 ± 1 kyr

(3.5 mm), 217 ± 0.8 kyr (12.5 mm), and 212 ± 0.7 kyr. The other subsamples of DR-20

represent a stalagmite that engulfed the stalactite. The age generated from the small stalagmite

(code 4S) was 199 ± 2.1 kyr at the top. Four U-series dates were measured from the uppermost

thin white layers: three at 3 mm (3A, 3B, and 3C) and one at 6 mm from the top surface (6).

The elevated U contents, between 2,450 and 4,820 ppb, and Th concentrations from

4,000 to 31,600 ppt (see Appendix A, Table A1), suggest a possible aragonitic composition

of the subsamples, but with high detrital component. The U-series age obtained for DR-D20-

6 was of 84 ± 0.3 kyr, whereas the DR-D20 (3A, 3B, and 3C) revealed discordant ages of 55

± 0.07, 64 ± 0.08, and 61 ± 0.3 kyr.

52

Figure 23. Cross section of sample DR-D20. Morphological components of the speleothem are indicated

and labeled. The original soda straw is shown in orange (dashed) and the stalactite-stalagmite contact in

black (line). Yellow: subsamples for U-series ages used in this study; Grey: subsamples that provided

unreliable U-series ages; Red: subsamples of whose ages were discarded and not included in the study.

Although all three calculated ages are very precise, the results are rather inconsistent

between them, which may suggest carbonate deposition and successive weathering anytime

between 85 and 60 ka.

4.6 U-series Systematics and Ages of Investigated Speleothems

To test the uranium systematics of all samples, an evolution curve (234U/238U ratio vs

230Th/238U) was generated using the Isoplot computer program (Ludwig, 2001) (Fig. 24).

The graphic displays the evolution patterns of samples having different initial δ234U

values. This allows to identify any obvious alteration of the calcite/aragonite minerals after

precipitation, and thus, an effective test supporting the reliability of the U-series ages. As

mentioned earlier, most samples with ages between 450 and 700 kyr cluster close to the secular

equilibrium (s.e.) line (red dashed in Fig. 24). A number of samples, specifically DIP-

53

flowstones, MIP-5-25, MIP-4A, MIP-3, and CF-13, plot on or just outside the curve range (in the

so-called ‘forbidden zone’), there has been some alteration.

Figure 24. Uranium-evolution curve for all speleothems in this study. Samples are color-coded

and described in the legend. In close-up, the samples clustering near the secular equilibrium line.

Ellipse dimensions express absolute 2σ errors.

In fact, none of these samples returned finite U-series ages, but when using the δ234U

model for those samples with low but non-zero δ234U measured values, some provided suggested

ages, although with high errors (Fig. 25).

54

Figure 25. Plot of all ages from all speleothems investigated in this study. Samples are site-based color-

coded and labeled. Errors bars are absolute 2σ. Shaded area represents the U-series limit range.

The samples approaching the U-series limit (~600 kyr) exhibit higher age uncertainties, as

well. This is due to the extremely low U content in the speleothems, which decreases age accuracy,

and to associated uncertainties. Nevertheless, the state-of-the-art analytical facilities employed at

UNM allow to precisely measure extremely low uranium concentrations in cave deposits, increasing

the suitability of the U-series technique to its upper limit (Asmerom et al., 2006).

Additionally, flowstones FE-1-A and FE-1-E plot slightly below the s.e. line, suggesting

post-depositional alteration of the uranium system (i.e., U loss). Sub-samples DR-D20-3 (A, B, C)

have been omitted from the evolution curve because of the post depositional weathering of its

aragonite layers as observed on the outer surface (Fig. 12C). Apart from these sub-samples, no other

evidence of open system behavior was detected in any of the investigated speleothems.

55

CHAPTER FIVE:

DISCUSSION

Here I discuss the results of our investigations suggesting new timeline insights on the

coastal caves genetic processes in the eastern Mallorca. In the first section, I describe in detail

the evolution of the different paleo-caves on a location-basis with major emphasis on the sea

level implications in the various speleogenetic phases. In the second part, I concentrate on two

of the speleothems examined, DR-D20 (Coves del Drac) and CF-11 (Cova de Cala Falcó),

discussing how they constrain the relative sea level between Marine Isotope Stages 7 and 5a and

their paleoenvironmental significance.

5.1 Mallorcan Coastal Karst and Relative Sea Level Changes over the Past 750 kyr

The long-term sea level variations in the western Mediterranean are mainly controlled by:

changes in size and distribution of ice sheets, which cause glacial isostatic (GIA) effects, thermal

expansion/contraction, dynamic topography (DT), and meteorological components (wind,

atmospheric pressure) (Gamis et al., 2012). The reconstruction of Mediterranean sea-level

(especially the western basin) using benthic and planktonic foraminifera δ18O is more complex

than at other sites (e.g., Red Sea) (Rohling et al., 2009). This is due to changes in salinity caused

by the Atlantic water input through the Strait of Gibraltar and the catchment basins of rivers

entering the Mediterranean Sea, which collect water from different climate regions (Rohling et

al., 2014). Given all these, sea level reconstructions based on oxygen isotopic composition of

carbonate microfossils in the central and western Mediterranean cannot be applied per se and

relative sea level (RSL) indicators such as speleothems (POS, submerged stalagmites with marine

56

biogenic encrustations, etc.), beach deposits, marine terraces, and shore platforms are

unequivocally essential to investigate past sea level changes in the central-western Mediterranean

region (Fornós et al., 2002a; Antonioli et al., 2004; Tuccimei et al., 2006; Dutton et al., 2009;

Dorale et al., 2010; Zazo et al., 2013; Lorscheid et al., 2017; Polyak et al., 2018). However, RSL

records for periods before ~250 kyr are rather poorly preserved and documented, most of which

have been described in the coastal caves of Mallorca Island (Dumitru et al., 2018). Alongside with

the POS, which are direct sea-level markers, vadose speleothems such as flowstones and

stalagmites are also used as indirect evidence of maximum past sea level stands (Onac et al., 2012;

Van Hengstum et al., 2015). These types of speleothems can precipitate under vadose conditions

in coastal caves by dripping/seeping of water percolating from surface into the cavities. In coastal

caves, episodes of flooding, mainly linked with sea level rise and/or increase of rainfall, cease

calcite (or aragonite) deposition, causing corrosion and weathering of the speleothems surface

(Harmon, 1978; Richards, 1994). Investigating such secondary cave deposits, relevant information

regarding coastal karst basins evolution, dynamics, and processes can be generated.

Our radiometric dating of speleothems exposed on the eastern coastline karst of Mallorca

suggest they formed between 650 and 450 ka when the relative sea level fluctuated from at least

two or more meters below present sea level (mbpsl) during MIS 17, 15, and 13 (interglacials) to

tens of meters below the present-day datum during cool stadials 16, 14, and 12 (Fig. 26). Since

RSL estimates of this period are rather scarce in the western Mediterranean, our results are

compared with two eustatic sea level (ESL) curves from Naish and Wilson (2008), and Spratt and

Lisiecki (2016), both generated from a stack of globally scattered benthic foraminiferal δ18O

records (Lisiecki and Raymo, 2005). The ESL curves (orange and blue in Fig. 26) suggest that

global mean sea level did not reached the present-day elevation, but peaked at ~5-7 mbpsl during

57

the MIS 17 and 15 interglacials. These reconstructions indirectly support our records from

Mallorca, although no precise estimates of relative sea level position can be generated for this

interval. Nevertheless, new benchmarks regarding the timing of distinct speleogenetic stages and

the alternation of phreatic-vadose conditions in the coastal karst of Mallorca are provided.

Figure 26. Ages of the speleothems from our study and eustatic sea level (ESL) reconstructions from

Naish and Wilson (2008) (orange curve), and Spratt and Lisiecki (2016) (blue curve). Speleothem

samples are color-coded. All age errors are absolute 2σ. Gray shaded box indicates the period of possible

breakdown and dismantling of the paleo-caves investigated in this study. Marine Isotopes Stages (MIS)

are also labeled.

In addition, I propose new insights into the littoral dynamics that occurred during the

precipitation of vadose speleothems and cave breakdown events as a consequence of multiple

cycles of sea level transgressions/regressions.

Models of coastal speleogenesis in the eogenetic karst of Mallorca were earlier presented

by Ginés and Ginés (1992) and Ginés (2000). More recent explorations of subaqueous passages in

58

caves such as Coves del Drac, Cova de sa Gleda, Cova Genovesa, and Cova des Pas de

Vallgornera brought new information on the speleogenetic processes, particularly those related

with the role of ceiling collapse in the caves’ development (Ginés et al., 2008; Gràcia et al., 2009,

2010; Ginés et al., 2014). By integrating our radiometric investigations with these previous

models, better frame the timescale of various speleogenetic phases (see Fig. 27) were provided,

which in turn allowed to place constraints on sea level changes during middle-to-late Pleistocene in

Mallorca. Sea level regression during late Messinian caused the subaerial exposure of the upper

Tortonian-early Messinian carbonate lithostratigraphic sequences (calcisiltites, calcarenites, and

reefal limestones). During this period, referred as the Messinian Salinity Crisis peak (MSC, 5.9-

5.3 Ma), the Mediterranean basin may have experienced a sea level drop of about 800-1200 m,

opening migration corridors that allowed terrestrial mammalian fauna (i.e., Myotragus sp.,

Hypolagus sp., etc.) to colonize the Balearic Islands (Mas et al., 2018). Before and after the MSC

peak, sea level may have oscillated at different rates and magnitudes. At times during these

intervals the mixtures of fresh and marine waters initiate the dissolution of the aragonitic

component of the reefal unit, generating proto-voids in the so-called mixing zone (Fig. 27-1).

Successive sea level rise during the early-middle Pliocene yielded to an upward movement of

mixing-corrosion zone intensifying the carbonate dissolution and further development of the

cavities (Fig. 27-2). U-Pb dating analyses of POS from Coves d’Artà (NE Mallorca) recorded sea

level positions between 23 and 32 mapsl from 4.5 to 3 Ma corroborating a rather long period of

high sea-stands in the western Mediterranean corresponding to the Pliocene Climatic Optimum

(4.4-4 Ma) and the Mid-Pliocene Warm Period (3.3-3 Ma) (Dumitru et al., 2018a). These new

Pliocene RSL records add relevant information to the sea level history of the Mediterranean basin

and support our interpretations of this coastal speleogenetic stage (Fig. 27-2) in Mallorca.

59

Figure 27. Conceptual model showing the different speleogenetic stages in the evolution of the coastal

caves in the eastern Mallorca [modified after Ginés and Ginés (1992) and Ginés et al. (2014)]:

1) Cave genesis in the mixing zone, 2) Sea level rise and further cave development, 3) Collapses

following sea level drop, 4) Vadose speleothems precipitation, 5) Partial collapse of ceiling and

cementation of breakdowns, and 6) Coastal erosion that exposes speleothems at surface

(See text for detailed explanation).

60

After this interval of relatively high sea-level stands, glacio- and hydro-eustatic-related

sea level variations continued to manifest throughout the late Pliocene-early Pleistocene in this

region of the Mediterranean as a response to ice-sheet waxing/waning (Raymo et al., 2006; Naish

et al., 2009; Pollard and DeConto, 2009). The proto-caves formed in the mixing zone

experienced periodic episodes of evolution under vadose (water free) and phreatic (flooded)

conditions. During sea-level high stands (phreatic stage), the cavities continued to increase their

volume due to further dissolution caused by mixing of marine and fresh water. As sea level

dropped, water abandoned the caves, and this triggered episodes of conspicuous breakdown

along the fractures and joints, generating larger caves (Fig. 27-3). In some cases, the limestone

blocks that accumulated on the cave floor were partially filled with fine-grained sand and clay

containing fossil remains. The sediments were transported underground by surficial runoff or via

percolating meteoric water during exceptional rainfall events. These periodic flooding and

collapse events may have persisted until middle-Pleistocene, creating caves of remarkable

dimensions.

As already mentioned, some middle Pleistocene sea level reconstructions indicate high

stands lower than today, and the caves along the eastern coast of the island may have

experienced long and substantial phases of development under vadose conditions. By applying a

conservative approach regarding the global mean sea level estimates to Mallorca (Fig. 26 the

seashore may have experienced significant regressions within ~100-200 m seawards linked

particularly to cool events (Fig. 27-4), which exposed extensive portions of the carbonate

platform to subaerial conditions. During this time, meteoric and sub-surface hydrological

processes (e.g., percolation, etc.) were the major factors in the development of caves by means of

extensive breakdown mechanisms (passage morphology, pattern, etc.), and noteworthy amounts

61

of flowstones and stalagmites were deposited (Fig. 27-4). Our study documents a synchronous

precipitation of vadose speleothems happening in different cave locations between ~650 and

~450 ka. This suggests periods with frequent rainfall events that helped maintaining a steady

drip water supply to caves and thus abundant speleothem deposition. The presence of a

conspicuous vegetation coverage may have increased the organic matter content in the soil, and

consequently more availability of CO2 in the percolating waters increased the dissolution of the

carbonate layers. This hypothesis is supported by our measurements that reveal moderately high

values of 232Th in the speleothem samples, indicating a strong contribution of non-radiogenic Th

from soil. Besides our radiometric investigations, dating of various speleothems from other

Mallorcan and western Mediterranean caves (Hodge, 2004), support our interpretations regarding

these vadose phases and RSL low stands during middle Pleistocene. However, the

geomorphological settings may have undergone significant changes throughout the middle and

upper Pleistocene, when sea level fluctuations caused recurrent flooding of the cavities that

ceased speleothems growth. Estimates of ESL variations, and RSL indicators from Spain and

Italy, suggest that sea level high stands occurred during the MIS 11, 9, 7, and 5 (Lisiecki and

Raymo, 2005; Tuccimei et al., 2006; Naish and Wilson, 2008; Dutton et al., 2009; Dorale et al.,

2010; Rohling et al., 2010; Zazo et al., 2013; Spratt and Lisiecki, 2016; Lorscheid et al., 2017;

Polyak et al., 2018; Stocchi et al., 2018). Throughout the interglacial periods the coastal caves of

Mallorca may have experienced complete or partial inundations, followed by intervals when

caves were water free (during glacials and stadials). Collapse events during sea-level low stands

may have triggered partial ceiling breakdowns, opening the cave to the surface (Fig. 27-5). This

process is in fact responsible for providing access to many of the littoral caves in this region, like

for instance Cova des Pas de Vallgornera (Ginés et al., 2014). The collapsed blocks along with

62

the speleothems accumulated inside the caves were subsequently cemented by carbonate sands

and red clay sediments containing fossil remains of terrestrial fauna. Sea level rise and

mechanico-chemical coastal erosion dissolution of the carbonate bedrock throughout the

Holocene induced the total collapse and destruction of the caves and removal of the breakdowns,

exposing the speleothems at surface (Fig. 27-6). A great variety of paleokarst structures are

nowadays visible along the eastern coast of Mallorca, showing distinct features and

morphologies that testify active coastal processes and RSL changes over the past 750 kyr. Here I

focus and describe in detail the following four sites: Cala Falcó, Sa Ferradura, Dimoni, and Na

Mitjana.

5.1.1 Cala Falcó Paleo-cave

The present-day morphology of Cala Falcó site is characterized by a narrow cala (bay)

and a small cave (Cova de Cala Falcó) opening at ~2 mapsl on the east side of a spectacular cliff

(Fig. 28A). Numerous breakdown blocks covering the bay and the first section of the cave,

along with abundant vadose speleothems outcropping near the current cave entrance (Fig. 28B-

C), attest episodic collapses and different speleogenetic phases occurred as direct consequence of

past sea level fluctuations.

The U-series dating of the exposed flowstones (yellow stars in Fig. 28C and

Appendix A, Table A1) indicate a period of vadose deposition between 650 and 450 ka

(Fig. 26), suggesting a relative sea level in the western Mediterranean basin lower than

today. Considering the ESL reconstructions produced by Naish and Wilson (2008) and

Spratt and Lisiecki (2016), a remarkable seashore regression during this interval at Cala

Falcó site may have exposed an extensive surface of the carbonate platform to meteoric

weathering and stream incision. In this coastal setting, a paleo-cave of large dimensions

63

must have been existed in Cala Falcó since flowstones and stalagmites were precipitated in

it, although no precise estimates regarding the size of this cavity can be put forward. The

outstanding sequences of vadose speleothems that nowadays outcrop in this site suggest that

the vadose conditions may have persisted for quite long time.

Figure 28. Images of Cala Falcó site: A) General view of the cala, cave entrance, and collapsed blocks,

B) Image of the paleokarst features outcropping on the cliff outside the cave, and C) Close up of the

paleo-flowstones and stalagmites. Yellow stars represent the sampling points, whereas the white lines

highlight the contour of the speleothems (photo credits: B.P. Onac and J. McIlrath).

64

In this coastal setting, a paleo-cave of large dimensions must have been existed in

Cala Falcó since flowstones and stalagmites were precipitated in it, although no precise

estimates regarding the size of this cavity can be put forward. The outstanding sequences of

vadose speleothems that are now exposed at this site suggest that the vadose conditions may

have persisted for quite long time.

As already described in our model (Fig. 27-5), consequent upper Pleistocene sea level

oscillations may have strongly influenced the morphology of the paleo-cave. Particularly,

during the last interglacial period (MIS 5e), RSL at Mallorca rose ~2 mapsl from 127 and

116 kyr, and GIA corrected POS records suggest that ESL rose up to ~6-7 mapsl early in

MIS 5e and dropped down to ~2 mapsl between at ~122 ka (Polyak et al., 2018).

Throughout this interval, the Cala Falcó paleo-cave was partially flooded. The sea level low

stands between MIS 5e and 5a and the drop following MIS 5a high stand (Dorale et al.,

2010) may have triggered a partial collapse of the present-day cave entrance and part of the

breakdown blocks were washed out throughout the Holocene exposing the vadose paleo-

speleothems that are visible today on the cliff.

5.1.2 Sa Ferradura and Dimoni Paleokarst

Sa Ferradura and Dimoni sites are two small littoral embayments graved in the upper

Miocene paleokarst funnel-like structures developed along the eastern coast of Mallorca, and

separated by a quite prominent promontory (Fig. 29A). The volumes of breccia deposits

existing at these two sites are rather conspicuous and document a long-lasting exposure of the

carbonate cliff to various coastal karst processes. Based on the morphology, abundance of

breakdowns, and widespread occurrences of exposed speleothems, different conditions must

have been existed in the past (Fig. 29B-D). In fact, the U-series ages obtained on flowstone

65

samples indicate an active interval of vadose speleothem deposition between ~650 and 470 ka

(Fig. 26). This clearly suggests that a series of well-developed littoral caves already existed at

Sa Ferradura and Dimoni prior to the deposition of speleothems, which may be somehow

related to the syngenetic paleokarst structures developed early during Messinian times (Fornós,

1999; Ginés et al., 2011). These ages are rather consistent with those from Cala Falcó

emphasizing our interpretations of lower RSL in Mallorca during this time period.

Considering the significant amounts of breccias and stacks of meter-size boulders, the paleo-

caves may have reached moderately large dimensions. Although I cannot estimate their exact

size, based on our geomorphological survey and data collection, a tentative reconstruction of

the extent of these paleo-caves is highlighted by the red dashed lines in Figure 29A.

According to our coastal karst timescale model, the paleo-caves may have developed and

expanded during the late Pliocene-early Pleistocene following sea level oscillations that

triggered periodic collapse events (see Fig. 27-3 and explanation therein for details).

Throughout the middle-Pleistocene when even the high sea level stands (MIS 17, 15, and 13)

were below present sea level, suitable conditions existed in these paleo-caves for the

precipitation of outstanding sequences of flowstones and stalagmites. The U-series dating

performed on the lower section of the Sa Ferradura flowstone (Fig. 29B) suggests that the

onset of vadose conditions may have occurred before ~650 ka, but later than ~1 Ma.

The sea level oscillations that occurred between ~400 and 150 ka (MIS 11-6) prompted

frequent changes from phreatic to vadose conditions in these caves and paleokarst voids, and

may have triggered the beginning of extensive collapses that affected many littoral caves.

During MIS 5e when the sea-level in Mallorca stood ~2 mapsl (Polyak et al., 2018), the caves

became once again partly or totally flooded.

66

Figure 29. Sa Ferradura and Dimoni sites: A) Aerial view of the two engulfments, B) Panoramic photograph of the Sa Ferradura site,

C) Flowstone sequence outcropping in between breccia deposits, D) Fossil flowstones and stalagmites exposed at Dimoni, and E) Funnel-like

embayment produced by the Messinian paleokarst structure and cave developed in the breccia at present sea level

(photo credits: B.P. Onac, J.J. Fornós, and J. Ginés).

67

The stability of drowned caves is normally higher then air-filled chambers since the

water provides support to walls and ceiling. At the end of MIS 5e (~116 ka), and especially

after the onset of MIS 4, the total collapse of the paleo-caves may have occurred as sea-level

dropped and water retreated from the cavities. This is when impressive limestone boulders and

broken speleothems accumulated on what was then the cave floor and were later cemented.

Ever since they formed, the chemical and mechanical actions of seawater during

Holocene, led to the development of several cavities within this breccia in the proximity of the

present-day seashore (Fig. 29E).

5.1.3 Na Mitjana Paleokarst

The paleokarst structures present at the Na Mitjana site (Fig. 30A) show evidence for a

more complex coastal karst evolution than at other sites previously described. One possible

reason for such complexity may be related with the longer time the Jurassic limestones were

exposed to both littoral and meteoric weathering.

Our speleothems’ data (see Appendix A, Table A1) reveal various possible scenarios and

speleogenetic processes that occurred at different times over the past ~1 Myr. The MIP-1

flowstone and stalagmite MIP-2 (Fig. 11B) indicate the presence of vadose conditions during the

period ~520-450 kyr, and are consistent with the speleothems from Cala Falcó, Sa Ferradura, and

Dimoni, further supporting our suggestion that the RSL in Mallorca was lower than today during

the above-mentioned interval. The elevations of the cave deposits (~7 mapsl), along with the

rough and steep morphology of the cliff in which they appear, imply the presence of once a

spacious littoral cave and/or different chambers interconnected by smaller voids.

Some of these little cavities that are today exposed in the paleokarst structure near the Na

Mitjana cave entrance (see B and C in Fig. 30A) present mammillary-like speleothems (MIP-3

68

and MIP-5) covering the walls and ceilings (Fig.30C and D). Their presence indicates intervals

of subaqueous conditions that may or may not be primarily related to sea level.

Figure 30. Na Mitjana site: A) Paleokarst structures and cavities (highlighted), B) Various speleothems

in one of the cavities, C) Close image of the sample MIP-5, D) Mammillary-like speleothem MIP-3,

and E) Flowstone MIP-4 (photo credits: B.P. Onac).

However, the speleothems are rather uniform and, unlike most POS deposits and the

mammillary calcite described by Polyak et al. (2008), were probably formed in a meteoric water

setting (vadose pool/geode). Thus, the sample MIP-3 that deposited ~580 ka (close to the MIS

15 peak) is in discordance with our overall vadose speleothems results and dismiss the sea level

implications in the underwater onset. The hypothetical precipitation of subaqueous calcite pool

coating in meteoric water is supported by the moderately high non-radiogenic Th (initial) values

measured in the samples (see Appendix A, Table A1). Moreover, these type of vadose pool

subaqueous calcite deposits, have been identified in other Mallorcan caves, i.e. Cova des Pas de

69

Vallgornera and Coves d’Artà. In the phases that preceded and followed these subaqueous

depositions (between ~590 and ~510 ka), precipitation of MIP-4 flowstone occurred (Fig. 30E).

According to our data on MIP-5 (see Appendix A, Table A1), favorable conditions for

calcite deposition in subaqueous conditions also existed in Mallorca at least during two other

periods: at ~1.1 Ma and between 296 and 186 ka. The latter interval (296-185 ka), which

includes the penultimate interglacial (MIS 7), was probably characterized by warmer and wetter

climate conditions that caused recurrent flooding of the voids.

5.2 Constraining Relative Sea Level Maximum in Mallorca during the MIS 7

Relative sea level during the penultimate interglacial (MIS 7) is poorly documented in the

central and western Mediterranean, and the few records available from different sites (i.e., E.

Spain, W. Italy, etc.) are rather contrasting. One of the main reasons for such mismatching data

is associated with the site-specific relationship between geological RSL indicators and tectonic

settings (e.g., uplifting, subsidence, relative stability, etc.) of the areas investigated. Marine

terraces that crop out in the eastern coast of Spain (La Marina-El Pinet) suggest the presence of

‘Senegalese’ warm fauna during MIS 7 in the western Mediterranean (Zazo et al., 2003). Open-

system U-series ages of ‘in situ’ coral colonies of Cladocora caespitosa and Strombus bubonius

shells from this site, yielded an indicative RSL estimate in the Spanish Mediterranean between

~1 mbpsl and ~1 mapsl (Goy et al., 2006; Zazo et al., 2013).

The POS record presented by Vesica et al. (2000) from Mallorca indicates a RSL

highstand of ~5 mapsl at ~233 ± 30 ka. Tuccimei et al. (2000a) documented RSL elevations

from POS in Grotta del Nettuno (NW Sardinia, Italy) at ~1.9 and ~2.5 mapsl at 234 ± 18 and 195

± 5 ka, respectively. Tuccimei et al. (2000b) also reported a RSL position at 8 mbpsl based on a

submerged POS in Cova den Passol (E. Mallorca) that precipitated ~201 ± 16 ka. Bard et al.

70

(2002) and Dutton et al. (2009) identified three major relative sea level highstands during the

MIS 7 from submerged stalagmites in Argentarola Cave (W. Italy). U-series measurements of

the cave calcite lamina alternating with serpulids layers, suggest that relative sea level in

Argentarola occurred above -18 m (relative to present datum) during MIS 7e (248-231 ka) and

MIS 7a (201-189 ka), but remained at or below 18 mbpsl during the MIS 7c (217-206 ka).

The speleothem DR-D20 from Coves del Drac shows different phases of precipitation,

revealed by an inner soda staw/stalactite section covered by layers of stalagmitic calcite and/or

aragonite (Fig. 12D). Similar features have been observed in many other coastal caves of

Mallorca, like for instance Cova des Pas de Vallgornera (Onac, pers. comm.). In the case of Llac

de les Delícies, steady dripping of water provided the suitable conditions for rapid soda

straw/stalactite formation between 225 and 212 ka.

Subsequently, the soda straw/stalactite reached the cave floor or top of the stalagmite

underlying the straw (ex, Fig. 31a), allowing for the growth of stalagmitic calcite layers to engulf

it. Thereafter, the stalactite broke or was corroded above the stalagmite (ex, Fig. 31b), and the

remaining stalactite was preserved within the stalagmite up until 198ka, when the stalagmite

growth ceased (Fig. 31c). Deposition of thin white aragonitic lamina on top of the stalagmite

(Fig. 12C) may have occurred at a later stage, but however, our U-series analyses on four

subsamples from this upper part have not returned reliable ages, indicating possible weathering

(see Appendix A, Table A1). This could suggest that recurrent cave flooding during the last

interglacial (MIS 5) and late Holocene, submerged the stalagmite and caused

corrosion/dissolution of the aragonitic layers.

71

Figure 31. Image of Llac de les Delícies (Coves del Drac) showing the hypothetical phases of DR-D20

precipitation: a) soda straws/stalactites reaching the cave floor, b) hanging portions of broken stalactites c)

stalagmite covering the broken part of a soda straw/stalactite at the cave floor (photo credits: J.J. Fornós).

The top of stalagmite DR-D20 is located at 1 mbpsl, and if growth was continuous from

225 to 198ka, this suggests that during MIS 7c and MIS 7a highstands, sea level did not peaked

above this depth. Figure 32 shows the DR-D20 records (red triangles) along with submerged

stalagmites from Argentarola and other POS-based RSL reconstructions previously published

(Vesica et al., 2000; Tuccimei et al., 2000a, b), which are compared with the RSL curve from

Waelbroeck et al. (2002) and the sea level reconstruction for the Red Sea (Siddal et al., 2003),

both generated using δ18O values from marine sediment (dashed and solid blue line,

respectively). Timing of DR-D20 stalagmite growth is consistent with both sea level curves but

show only a partial agreement with the RSL estimates based on the Argentarola samples.

72

Figure 32. MC-ICP-MS U-series ages for DR-D20 speleothem (red triangle) are compared with the

Argentarola stalagmites (Bard et al., 2002; Dutton et al., 2009) (orange, purple and black squares), POS

TIMS and α-spectrometry ages from Grotta del Nettuno, Sardinia (Tuccimei et al., 2000a), and POS

α-spectrometry ages from Cove den Passol and Na Mitjana Cave. Mallorca (Tuccimei et al., 2000b;

Vesica et al., 2000). RSL curve (Waelbroeck et al., 2002) and Red Sea level reconstruction (Siddall

et al., 2003) are also shown (solid and dashed blue lines). All errors are absolute 2σ, except for

α-spectrometry ages from Tuccimei et al. (2000a, b), which are 1σ).

Discrepancies between our data and the POS records from Mallorca and Sardinia also are

noted. This could be explain in four ways: 1) problems associated with the MIS 7 POS ages that

show rather large uncertainties since they were obtained using U-series α-spectrometry

technique; 2) post-depositional diagenetic and/or weathering processes affected POS and hence

altered the true age; 3) POS-sea level elevations were not corrected for GIA and DT effects; and

4) different uplift rates between Mallorca and Sardinia islands (Polyak et al., 2018).

73

Although the GIA-corrected POS-based MIS 5e sea level record of Polyak et al. (2018)

compare well with the Waelbroeck-based GIA model, for the penultimate glacial/interglacial

cycle, the latter is rather contrasting with some of the Mediterranean RSL records shown in

Figure 32. Particularly, when comparing this curve with the sea level estimates from

Argentarola, many of the stalagmite ages plot below Waelbroeck-inferred sea surface. In

contrast, the comparison between the sea level reconstructions for Red Sea (Siddall et al., 2003)

and the RSL records from central and western Mediterranean are in better agreement. In a more

general sense, mismatching between benthic ESL reconstructions and RSL geological archives is

related to changes in deep water temperature, temporal variations of the melted ice isotopic

composition, as well as salinity and local δ18O benthic variations not related to sea level changes

(Chappel and Shackelton, 1986; Shackleton, 1987; Rohling and Bigg, 1998; Clarke et al., 2002).

In contrast, the geological RSL indicators can be affected by different local and regional

uplift/subsidence movements. In any case, due to such potential complications of the sea level

records, over-/underestimates of global mean and relative sea level variations can occur. In our

dataset, these discrepancies are not present, hence it can be suggested that DR-D20 provides

reliable constraints for a maximum relative sea level below 1 m relative to present during the

penultimate interglacial in Mallorca within the limits of the U-series absolute ages and errors.

However, precisely RSL elevations were determined due to the heterogeneity of the other direct

RSL markers. Thus, further investigations on POS and additional sea level records are needed to

make precise estimates of RSL elevations in the western Mediterranean basin.

5.3 MIS 5b/5a Transition and MIS 5a Sea Level Highstand in Mallorca

The last interglacial is one of the most studied and documented interval of the Quaternary

climatic history since it is often used as a potential analogue for future scenarios of sea level rise

74

(Kopp et al., 2009; Dutton and Lambeck, 2012; Dutton et al., 2015). However, during the MIS 5

(substages a-e) the global sea level fluctuated significantly in response to waxing/waning of the

ice sheets, and the magnitude and timing of these oscillations are still a matter of debate. In

particular, the MIS 5a sea level highstand has been recorded from different locations around the

world ranging between -30 to +3 m compared to present sea level ( Gallup et al., 1994; Lambeck

et al., 2001; Muhs et al., 2002; Wehmiller et al., 2004).

In the western Mediterranean region, investigations of POS from various caves of

southeastern Mallorca documented a RSL highstand at ~1-1.5 mapsl between 82 and 80 ka

(Tuccimei et al., 2006; Dorale et al., 2010). The CF-11 stalagmite from Cova de Cala Falcó was

recovered at ~1 mapsl, and the U-series measurements indicate it grew from 85.7 to 83.7 ka,

corresponding to the transition from substage MIS 5b to MIS 5a. During this interval, suitable

conditions of steady drip water supply must have existed in the cave, allowing the precipitation

of CF-11 stalagmite (Fig. 33A). Tuccimei et al. (2006) also documented a POS-based RSL in

Mallorca lower than ~15 mbpsl between 86.5 and 85.5 ka, which corroborates the existence of

vadose conditions in Cova de Cala Falcó at the time when CF-11 began to grow. The subsequent

rapid sea level rise between 83.7 and 80 ka caused the partial flooding of the cave that

submerged the stalagmite ceasing calcite precipitation (Fig. 33B). As described by Dorale et al.

(2010), sea level in Mallorca persisted at ~1.5 mapsl for a relatively short time (~2 kyr) and

rapidly dropped to -15 m (relative to present sea level) by ~78.6 ± 0.8 ka at the onset of MIS 4

cold stadial (Tuccimei et al., 2006). This sudden sea level fall exposed the CF-11 stalagmite to

subaerial processes that caused partial corrosion/weathering of the outer surface (Fig. 33C). In

this interpretation, dripping and continued precipitation of the stalagmite ceased.

75

High-resolution U-series analysis of the stalagmite CFF-03-03-1 from Cova de Cala

Falcó recorded a substantial growth hiatus from ca. 85 to 80 ka, coincident with the MIS 5a

marine transgression (Hodge et al., 2008). However, CF-11 stalagmite suggests that until ~83.7

ka the cave remained water-free, and the flooding took place after (~83 ka) in the early part of

the MIS 5a.

Figure 33. Conceptual drawing of CF-11 stalagmite evolution and sea level implications: A) stalagmite precipitation in vadose conditions during sea level lowstand, B) cave partial flooding

during MIS 5a sea level higstand cause the cessation of the stalagmite growth, and C) sea level fall

and subaerially exposure of the stalagmite.

In contrast, the CAM-1 stalagmite from Cova de Campanet (NW Mallorca) shows a

continuous precipitation throughout the MIS 5b and 5a interval (Dumitru et al., 2018b). This

confirms that in inland caves located at higher elevations, optimal conditions continued to exist

76

for speleothem deposition, whereas in coastal caves (e.g., Cala Falcó), the sea-level rise

associated with MIS 5a highstand caused a temporal or complete cessation of stalagmite growth.

Figure 34. Comparison of Cala Falcó (CF-11) stalagmite oxygen composition with other local and

global records: a) NGRIP (North Greenland Ice Core Project members, 2004; Rasmussen et al., 2014),

b) δ18O records of CF-11, c) Campanet stalagmite δ18O values (Dumitru et al., 2018b), d) δ18O of Cala

Falcó (CFF-03-03-1) stalagmite (Hodge et al., 2008), and e) Sea level reconstructions for Mallorca during

MIS 5a, based on phreatic overgrowths on speleothems (Tuccimei et al., 2006; Dorale et al., 2010) (blue

and red circles, respectively). U-series ages of the POS (2σ absolute errors) and elevations related to

present sea level are shown. MIS boundaries are also displayed. MIS 5a highstand is shown in light blue.

An overall positive trend of the δ18O values is observed in Greenland ice core, CF-11,

and Campanet between 85.5 and 84.9 ka (Fig. 34 a-c), suggesting for Mallorca a period of

77

ameliorated (milder) climate conditions when cave drip waters mainly originated from 18O-

enriched Mediterranean-sourced moisture (Fig. 34b and c).

Thereafter, CF-11 δ18O values show a sudden decrease coinciding with a century-long

cold event (see arrow in Fig. 34a and b) recorded towards the end of the Greenland Stadial 22

(Rasmussen et al., 2014). The complex δ18O profile reported by Hodge (2004) for CFF-03-03-1

stalagmite, not only vary between 0 and -10‰ (double than our range), but also shows a large

number of rather regular zigzagging back and forth of the data compare to CF-11 data (Fig. 34d).

Because of this direct comparison between the two records could not be made. It could be that

this fluctuating pattern prevented Hodge et al. (2008) to use the δ18O values in their study.

The δ13C values in the Cala Falcó stalagmites (CF-11 and CFF-03-03-1) are rather

consistent and show a small increase between 85 and 84.6 ka. Slightly different variation trends

are observed at 84.6 ka and from 84.4 to 84.2 ka, which could be related with distinct percolating

water paths and/or dripping rates in the different parts of the cave. When comparing the δ13C

values for the two Cala Falcó stalagmites with the record from Campanet, a general positive

trend can observed until ~84.7 ka (Fig. 35).

Afterwards, a step-wise δ13C decrease is observed in Cala Falcó (Fig. 35a and b), whereas

the records from Campanet show a slow increase (Fig.35c). During the MIS 5b/5a transition

(~84.2-83.8 ka), the δ13C values from CF-11 and CAM-1 present opposite trends. One

explanation for these contrasting δ13C trends could be the difference in the amount of biogenic

CO2 production/supply and/or the C3/C4 plant ratio above the caves. Nevertheless, no C4 plant

species have been documented in Mallorca (Bolòs, 1996), and speleothem C3-sourced δ13C

values are generally more negative (between -14 and -6‰; Rudzka et al., 2011) than those

observed in the Mallorcan stalagmites.

78

Figure 35. δ13C comparison between: a) Cala Falcó (CFF-03-03-1) stalagmite (Hodge et al., 2008),

b) Cala Falcó (CF-11) stalagmite, and c) Campanet speleothem (Dumitru et al., 2018b), MIS 5b/5a

boundary is shown.

Therefore, I suggest that the differences in the δ13C values between Cala Falcó and

Campanet stalagmites are related to more local imprints, such as thickness of the soil horizon and

biological activity therein, prior calcite precipitation in the epikarst zone (Fairchild and Treble,

2009), and possibly the storage capacity above the two caves, which regulates the amount of

percolating water and consequently the cave drip rate.

79

CHAPTER SIX:

CONCLUSIONS

Our findings yield new geochronological insights regarding the speleogenetic evolution

of paleo-coastal caves and sea level variations in Mallorca over the last 750 kyr. Specifically, I

provide additional timeframe evidences for the previous coastal karst models proposed by Ginés

and Ginés (1992) and Ginés et al. (2014), and used these data to constraint sea level positions

during the different speleogenetic stages. Our radiometric analyses of paleo-speleothems

indicate long-lasting phases of vadose conditions along the present-day eastern coastline between

650 and 450 ka, suggesting that relative sea level positions during the MIS 17, 15, and 13

highstands may not have exceeded ~2 mapsl. The occurrence of conspicuous sequences of

flowstones and stalagmites at the various investigated paleokarst sites (e.g., Cala Falcó, Sa

Ferradura, Dimoni, and Na Mitjana) also hint to the presence of caves having remarkable

dimensions, and at the same time, indicate that suitable conditions (warm and wet) existed

throughout this interval. However, no precise assessments concerning the paleo-caves extension

can be formulated from the data available. Sea level fluctuations during the glacial/interglacial

cycles caused caves to experience periodic phreatic and vadose phases that affected the structure

and morphology of these paleo-caves. I suggest that recurrent floodings occurred during MIS

11, 9, 7, and 5e interglacials, followed by periodic drawdowns during the glacial stages of MIS

10, 8, 6, and 4. Throughout these cool stadials, the paleo-caves were affected by partial ceiling

breakdowns and subsequently cementation of the collapsed blocks and speleothems. Since the

last interglacial sea level highstands and throughout the Holocene, littoral erosion and dissolution

80

processes caused partial/total collapse of the paleo-caves, exposing a variety of vadose

speleothems at the surface.

Records of sea level maximum are documented in the stalagmite DR-D20 from Coves del

Drac. The speleothem precipitated between 225 and 198 ka, suggesting that the relative sea level

in Mallorca did not reach 1.1 mbpsl during the MIS 7a and MIS 7c highstands. The data are

partly in agreement with distinct RSL records from the central Mediterranean and fairly

consistent with other sea level reconstructions (global and Red Sea), although show various

degrees of discordances with some POS-based RSL records from Mallorca and Sardinia islands.

Because of the scarcity of RSL records for MIS 7 in the western and central Mediterranean, and

the rather heterogeneity of the data available, no precise sea level estimates are generated.

Further investigations are necessary in order to precisely assess the RSL positions during the

penultimate interglacial.

Additionally, the U-series ages of the CF-11 stalagmite from Cova de Cala Falcó indicate

an interval of precipitation between 85.7 and 83.7 ka, coinciding with the latter phase of the

stadial MIS 5b and the onset of MIS 5a (MIS 5b/5a transition). High-resolution isotopic

comparisons between CF-11 and other local and global proxy records suggest that mild climate

conditions were present in Mallorca from 85.5 to 84.9 ka with Mediterranean-sourced

precipitations that supplied steady dripping water into the cave. At the beginning of the MIS 5a,

Cala Falcó stalagmites ceased to grow, whereas the inland speleothem from NW Mallorca

continued to precipitate throughout all the interval. POS-based RSL reconstructions for

Mallorca document that between 85 and 82 sea level rapidly arose to ~1.5 mapsl, confirming that

partial flooding of the cave occurred soon after 83.7 ka, submerging the stalagmite, which

stopped growing.

81

In conclusion, our records from Mallorca further our understanding of coastal karst

processes and their relationship with past sea level changes in the western Mediterranean,

indicating the suitable utility of vadose speleothems (with their limitations) as records of past sea

level variations. Our study suggests that the application of such records along with other direct

sea level markers (e.g., POS), will constitute extremely powerful tools for future sea level

investigations.

82

REFERENCES

Andersen, M.B., Stirling, C.H., Potter, E.K., Halliday, A.N. (2004). Toward epsilon levels of

measurement precision on 234U/238U by using MC-ICPMS. International Journal of Mass

Spectrometry, 237 (2-3): 107-118.

Antonioli, F., Bard, E., Potter, E.K., Silenzi, S., Improta, S. (2004). 215-ka History of sea-level

oscillations from marine and continental layers in Argentarola Cave speleothems (Italy). Global

and Planetary Change, 43 (1-2): 57-78.

Asmerom, Y., Polyak, V.J.., Schwieters, J., Bouman, C. (2006). Routine high-precision U–Th

isotope analyses for paleoclimate chronology. Geochimica et Cosmochimica Acta, 70 (18): A24.

Balaguer, P. (2005). Tipus i evolució de les costes rocoses de Mallorca. Unpublished PhD

Thesis. Palma de Mallorca: Universitat de les Illes Balears, 373 p.

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 (3-4): 135-146.

Barnes, J.W., Lang, E.J., Potratz, H.A. (1956). Ratio of ionium to uranium in coral limestone.

Science, 124: 175-176.

Baskaran, M., Krishnamurthy, R.V. (1993). Speleothems as proxy for the carbon isotope

composition of atmospheric CO2. Geophysical Research Letters, 20: 2905-2908.

Becquerel, H. (1986). Émission de radiation nouvelles par l'uranium métallique. Comptes

Rendus, 122: 1086.

Boehnke, P., Harrison, T.M. (2014). A meta-analysis of geochronologically relevant half-lives:

what’s the best decay constant? International Geology Review, 56: 905-914.

Bögli, A. (1980). Karst hydrology and physical speleology. Berlin: Springer-Verlag, 284 p.

Bolòs, O. (1996). La vegetació de les Illes Balears. Barcelona: Institut d'Estudis Catalans, 269 p.

Chappell, J., Shackleton, N. (1986). Oxygen isotopes and sea level. Nature, 324(6093): 137-140.

Chen, J.H., Edwards, R.L., Wasserburg, G.J. (1986). 238U, 234U and 232Th in seawater. Earth and

Planetary Science Letters, 80 (3-4): 241-251.

Cheng, H., Edwards, R.L., Hoff, J., Gallup, C.D., Richards, D.A., Asmerom, Y. (2000). The

half-lives of uranium-234 and thorium-230. Chemical Geology, 169 (1-2): 17-33.

83

Cheng, H., Edwards, R.L., Shen, C.C., Polyak, V.J., Asmerom, Y., Woodhead, J., Hellstrom, J.,

Wang, Y., Kong, X., Spötl, C., Wang, X., Alexander Jr., C.E. (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:

82-91.

Cherdyntsev, V.V., Kazache. I., Kuzmina, Y.A. (1965). Dating of Pleistocene carbonate

formations by Thorium and Uranium isotopes. Geochemistry International, 2 (5): 794-798.

Clark, P.U., Shakun, J.D., Marcott, S.A., Mix, A.C., Eby, M., Kulp, S., Schrag, D. P., et al.

(2016). Consequences of twenty-first-century policy for multi-millennial climate and sea-level

change. Nature Climate Change, 6 (4): 360-369.

Clarke, G.K., Marshall, S.J. (2002). Isotopic balance of the Greenland Ice Sheet: modelled

concentrations of water isotopes from 30,000 BP to present. Quaternary Science Reviews, 21 (1-3):

419-430.

Clemmensen, L.B., Fornós, J.J., Rodríguez-Perea, A. (1997). Morphology and architecture of a

late Pleistocene cliff -front dune, Mallorca, Western Mediterranean, Terra Nova. 9: 251-254.

Clemmensen, L., Lisborg, T., Fornós, J., Bromley, R.G. (2001). Cliff-front aeolian and colluvial

deposits, Mallorca, Western Mediterranean: A record of climatic and environmental change

during the last glacial period. Bulletin of the Geological Society of Denmark, 48: 217-232.

Cliff, R.A., Spötl, C., Magini, A., (2010). U\Pb dating of speleothems from Spannagel Cave,

Austrian Alps: a high-resolution comparison with U-series ages. Quaternary Geochronology, 5:

452-458.

Cole, J.M., Rasbury, E.T., Montañez, I.P., Pedone, V.A., Lanzirotti, A., Hanson, G.N. (2004).

Petrographic and trace element analysis of uranium‐rich tufa calcite, middle Miocene Barstow

Formation, California, USA, Sedimentology, 51: 433-453.

Curie, M.S., (1898) Rayons emis par les composes de l’uranium et du thorium, Comptes Rendus.

126: 1101.

Decker, D.D., Polyak, V.J., Asmerom, Y., Lachniet, M.S. (2018). U–Pb dating of cave spar: A

new shallow crust landscape evolution tool. Tectonics, 37: 208-223.

DeConto, R.M., Pollard, D. (2016). Contribution of Antarctica to past and future sea-level rise.

Nature, 531: 591-597.

Dickin, A.P. (2005). Radiogenic Isotope Geology (2nd edition). Cambridge: Cambridge

University Press, 508 p.

Dorale, J.A., Edwards, R.L., Ito, E., Gonzalez, L.A. (1998). Climate and vegetation history of the

midcontinent from 75 to 25 ka: a speleothem record from Crevice Cave, Missouri, USA. Science,

282: 1871-1874.

84

Dorale, J.A., Edwards, R.L., Onac, B.P. (2002). Stable Isotopes as Environmental Indicators in

Speleothems. In: Yuan, D., Zhang, C. (eds.), Karst Processes and the Carbon Cycle. Beijing:

Final Report of IGCP 379, Geological Publishing House, pp. 107-120.

Dorale, J.A., Edwards, R.L., Alexander, E.C., Shen, C.C., Richards, D.A., Cheng, H. (2004)

Uranium-Series Dating of Speleothems: Current Techniques, Limits, & Applications. In:

Sasowsky, I.D., Mylroie, J. (eds). Studies of Cave Sediments. Boston: Springer, pp. 177-197.

Dorale, J.A., Onac, B.P., Fornós, J.J., Ginés, J., Ginés, A., Tuccimei, P., Peate, D.W. (2010).

Sea- level 81,000 years ago in Mallorca. Science, 327: 860-863.

Dumitru, O.A., Onac, B.P., Polyak, V.J., Austermann, J., Fornos, J.J., Gines J., Gines A.,

Asmerom, Y. (2018a). A U-Pb Dated Glacial Isostasy corrected Pliocene sea-level record in

Mallorca (western Mediterranean) based on phreatic overgrowths on speleothems. Geological

Society of America, Abstracts with Programs, 50 (6): 222-7.

Dumitru, O.A., Onac, B.P., Polyak, V.J., Wynn, J.G., Asmerom, Y., Fornós, J.J. (2018b). Climate

variability in the western Mediterranean between 121 and 67 ka derived from a Mallorcan

speleothem record. Palaeogeography, Palaeoclimatology, Palaeoecology, 506: 128-138.

Dutton, A., Bard, E., Antonioli, F., Esat, T.M., Lambeck, K., McCulloch, M.T. (2009). Phasing

and amplitude of sea-level and climate change during the penultimate interglacial. Nature

Geoscience, 2 (5): 355-359.

Dutton, A., Lambeck, K. (2012). Ice volume and sea level during the last interglacial. Science,

337: 216-219.

Dutton, A., Carlson, A. E., Long, A., Milne, G. A., Clark, P. U., DeConto, R., Horton, B.P.,

Rahmstorf, S., Raymo, M. E. (2015). Sea-level rise due to polar ice-sheet mass loss during past

warm periods. Science, 349: 153-aaa4019.

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 y. Earth and Planetary Science Letters, 81:

175-192.

Esteban, M. (1979). Significance of the Upper Miocene reefs of the Western Mediterranean.

Paleogeography, Paleoclimatology, Paleoecology, 29: 169-188.

Faure, G., Mensing, T.M. (2005) Isotope Principle and Applications (3rd edition). Hoboken: John

Wiley & Sons, 928 p.

Fairchild, I.J., Smith, C.L., Baker, A., Fuller, L., Spötl, C., Mattey, D., McDermott, F. (2006).

Modification and preservation of environmental signals in speleothems. Earth-Science Reviews,

75 (1-4): 105-153.

Fairchild, I.J., Treble, P.C. (2009). Trace elements in speleothems as recorders of environmental

change. Quaternary Science Reviews, 28 (5-6): 449-468.

Fairchild, I.J., Baker, A. (2012). Speleothem science: from process to past environments.

Chichester: John Wiley & Sons, 432 p.

85

Ford, D.C., Thompson, P.L., Schwarcz, H.P., (1972). Dating cave calcite by the uranium

disequilibrium method; some preliminary results from Crowsnest Pass, Alberta, International

Geography, 22: 21.

Fornós, J.J., Pomar, L. (1983a). Mioceno superior de Mallorca: Unidad Calizas de Santanyí

(“ComplejoTerminal”). In: Pomar, L., Obrador, A., Fornós, J.J., Rodriguez-Perea, A. (eds.), El

Terciario de las Baleares (Mallorca-Menorca). Palma de Mallorca: Guía de excursiones, 10th

Congreso Nacional de Sedimentología, Institut d’Estudis Balearics and Universidad de Illes

Balears, 177-206.

Fornós J.J., Pomar L. (1983b). El complejo de manglar del Mioceno terminal de la isla de

Mallorca. Bulletí de la Societat d’Història Natural de les Illes Balears, 26: 207-228.

Fornós J.J., Pomar L. (1984). Facies, ambientes y secuencias de plataforma carbonatada somera

(Formación de Calizas de Santanyí), en el Mioceno Terminal de Mallorca (Islas Baleares),

Publicaciones de Geología. 20: 319-330.

Fornós, J.J., Gelabert, B. (1995). Litologia i tectònica del carst de Mallorca/ Lithology and

tectonics of the Majorcan karst. In A. Ginés, J. Ginés (eds.), El carst i les coves de

Mallorca/Karst and caves in Mallorca. Palma de Mallorca: Endins/Monografies de la Societat

d’Historia Natural de les Balears, 20 (3): 27-43.

Fornós, J.J., Ginés, J., Ginés, A., Rodríguez-Perea, A. (1995). Paleokarst in Mallorca.

Endins/Monografies de la Societat. d’Historia Natural de les Balears, 20 (3): 113-124.

Fornós. J.J. (1999). Karst collapse phenomena in the Upper Miocene of Mallorca (Balearic

Islands, Western Mediterranean). Acta Geologica Hungarica, 42 (2): 237-250.

Fornós, J.J., Gelabert, B., Ginés, A., Ginés, J., Tuccimei, P., Vesica, P. (2002a). Phreatic

overgrowths on speleothems: a useful tool in structural geology in littoral karstic landscapes. The

example of eastern Mallorca (Balearic Islands). Geodinamica Acta, 15 (2): 113-125.

Fornós, J.J., Pomar, L., Ramos-Guerrero, E. (2002b). Balearic Islands, Tertiary. In Gibbons, W.,

Moreno, T. (eds.), The geology of Spain. London: The Geological Society, pp. 327-334, 405-406.

Fornós, J.J., Clemmensen, L.B., Gómez-Pujol, L., Ginés, A., Ginés, J. (2012). Pleistocene

eolianites and low sea levels. 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’Historia Natural de les Balears, 18: 85-110.

Fornós J.J., Ginés À., Ginés J., Gràcia F., Merino A. (2019). Littoral Endokarst from Mallorca

Island (Western Mediterranean): Dynamic processes, Sediments and Management. In: Morales J.

(eds), The Spanish Coastal Systems. Cham: Springer Nature, pp. 143-166.

Gallup, C.D., Edwards, R.L., Johnson, R.G. (1994). The timing of high sea levels over the past

200,000 years. Science, 263: 796-800.

Gascoyne, M. (1992). Palaeoclimate determination from cave calcite deposits. Quaternary

Science Reviews, 11: 609-632.

86

Gelabert-Ferrer, B. (1998). La estructura geológica de la mitad occidental de la isla de

Mallorca. Madrid: Memorias del Instituto Tecnológico Geominero de España, 129 p.

Giménez, J., Gelabert, B. (2002). Recent Tectonic Activity Analysis of Mallorca Island. EGS

General Assembly Conference Abstracts, 27: 390-394.

Ginés, A.; Ginés, J., Pomar, L. (1981). Phreatic speleothems in coastal caves of Majorca (Spain)

as indicators of Mediterranean Pleistocene paleolevels. Bowling Green: Proc. 8th International

Congress of Speleology, 2: 533-536.

Ginés, A., Ginés, J. (1987). Características espeleológicas del karst de Mallorca. Endins, 13: 3-19.

Gines A., Gines J. (1992). Las Coves del Drac (Manacor, Mallorca). Apuntes historicos y

espeleogeneticos. Endins, 17-18: 5-20.

Ginés, A., Ginés, J. (1995). The exokarstic landforms of Mallorca Island. In: Ginés, A., Ginés, J

(eds.). Karst and Caves in Mallorca. Endins/Monografies de la Societat d’Historia Natural de

les Balears, 20 (3): 56-70.

Ginés, A. (1999). Agriculture, grazing and land use changes at the Serra de Tramuntana karstic

mountains. International Journal of Speleology, 28 B (1/4): 5-14.

Ginés, A., Ginés, J. (2007). Eogenetic karst, glacioeustatic cave pools and anchialine

environments on Mallorca Island: A discussion of coastal speleogenesis. International Journal of

Speleology, 36 (2): 57-67.

Gines J., Gines A., Fornos J.J., Gràcia F., Merino A., (2008). Noves observacions sobre

l’espeleogenesi en el Migjorn de Mallorca: els condicionants litologics en alguns grans sistemes

subterranis litorals. Endins, 32: 49-79.

Ginés, A., Ginés, J. (2011). Classificació morfogenètica de les cavitats càrstiques de les Illes

Balears. Endins/Monografies de la Societat d’Historia Natural de les Balears, 17 (35): 85-102.

Ginés, A., Ginés, J., Fornós, J. J., Bover, P., Gómez-Pujol, L., Gràcia, F., Merino, A., Vicens, D.

(2012). An introduction to the Quaternary of Mallorca. 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’Historia Natural de les Balears, 18: 13-53.

Ginés, J. (1995). L’endocarst de Mallorca: els mecanismes espeleogenètics / Mallorca’s

endokarst: the speleogenetic mechanisms. Endins/Monografies de la Societat. d’Historia Natural

de les Balears, 20 (3): 71-86.

Ginés, J. (2000). El karst litoral en el levante de Mallorca: una aproximación al conocimiento de

su morfogénesis y cronología. Unpublished PhD Thesis. Palma de Mallorca: Departament de

Ciències de la Terra, Universitat de les Illes Balears, p. 595.

Ginés, J., Ginés, A., Fornós, J.J., Merino, A., Gràcia, F. (2009). On the role of hypogene

speleogenesis in shaping the coastal endokarst of southern Mallorca (western Mediterranean). In

A.B. Klimchouk, D.C. Ford (eds.), Hypogene speleogenesis and karst hydrogeology of artesian

basins. Simferopol: Ukrainian Institute of Speleology and Karstology, Special Paper 1: 91-99.

87

Ginés, J., Ginés, A., Fornós, J.J. (2011). Dades sobre paleokarst i espeleocronología de les illes

balears. Endins, 35 (17): 213-226.

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’Historia

Natural de les Balears, 18: 111-146.

Ginés, J., Fornós, J.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): 105-124.

Ginés, J., Fornós, J.J., Gràcia, F., Merino, A., Onac, B.P., Ginés, A. (2017). Hypogene Imprints

in Coastal Karst Caves from Mallorca Island (Western Mediterranean): Morphological Features

and Speleogenetic Approach. In: Klimchouk, A., Palmer, N.A., De Waele, J., Auler, A.S, Audra,

P. (eds.), Hypogene Karst Regions and Caves of the World. Cham: Springer International

Publishing, pp. 99-112.

Gómez-Pujol, L., Gelabert, B., Fornós, J.J., Pardo-Pascual, J.E., Rosselló, V.M., Segura, F.S.,

Onac, B.P. (2013). Structural control on the presence and character of calas: Observations from

Balearic Islands limestone rock coast macroforms. Geomorphology, 194: 1-15.

Gomis, D., Tsimplis, M., Marcos, M., Fenoglio-Marc, L., Pérez, B., Raicich, F., Vilibić, I.,

Wöppelmann, G., Monserrat, S. (2012). Mediterranean Sea-level variability and trends. In:

Lionello, P. (ed.), The climate of the Mediterranean region: From the past to the future. London:

Elsevier, pp. 257-299.

Gornitz, V. (2005). Eustasy. In: Schwartz, M.L. (ed.), Encyclopedia of Coastal Science.

Encyclopedia of Earth Sciences Series. Springer, pp. 439-442.

Goy, J.L., Hillaire-Marcel, C.I., Zazo, C., Ghaleb, B., Dabrio, C.J., Bardají, T., Civis, J., Preda,

M., Yébenes, A., Forte, A. M. (2006). Further evidence for a relatively high sea level during the

penultimate interglacial: open-system U-series ages from La Marina (Alicante, East Spain).

Geodinamica Acta, 19 (6): 409-426.

Gràcia, F., Fornos J.J. & Merino A. (2009) Corrosion patterns related to meteoric-marine mixing

zone in coastal cave systems of Mallorca Island (Western Mediterranean). In: White W.B. (eds.),

Proceedings of the 15th International Congress of Speleology, Kerrville, TX, 1: 496-500.

Gràcia, F., Clamor B., Gamundi P., Fornos J.J. (2010). El sistema de cavitats Gleda - Camp des

Pou (Manacor, Mallorca). Endins, 34: 35-68.

Guijarro, J.A. (1995). Aspectes bioclimatics del karst de Mallorca / Bioclimatic aspects of karst

in Mallorca. In: Ginés, A., Ginés, J. (eds.) El carst i les coves de Mallorca / Karst and caves in

Mallorca. Palma de Mallorca: Endins/Monografies de la Societat. d’Historia Natural de les

Balears, 20 (3): 17-26.

88

Harmon, R.S., Thompson, P., Schwarcz, H.P., Ford, D.C. (1978). Late Pleistocene Paleoclimates

of North America as Inferred from Stable Isotope Studies of Speleothems. Quaternary Research,

9 (1): 54-70.

Harmon, R.S., Land, L.S., Mitterer, R.M., Garett, P., Schwarcz, H.P., Larson, G.J. (1981).

Bermuda sea level during the last interglacial. Nature, 289: 481-483.

Hartmann, A., Baker, A. (2017). Modelling karst vadose zone hydrology and its relevance for

paleoclimate reconstruction. Earth-Science Reviews, 172: 178-192.

Hellstrom, J. (2006). U-Th dating of speleothems with high initial 230 Th using stratigraphical

constraint. Quaternary Geochronology, 1: 289-295.

Hendy, C.H., Wilson, A.T. (1968). Palaeoclimatic data from speleothems. Nature, 219: 48-51.

Hendy, C.H. (1971). The isotopic geochemistry of speleothems-I. The calculation of the effects

of different modes of formation on the isotopic composition of speleothems and their

applicability as palaeoclimatic indicators. Geochimica et Cosmochimica Acta, 35 (8): 801-824.

Hill, C., Forti, P. (1997). Cave Minerals of the World (2nd edition). Huntsville: National

Speleological Society, 463 p.

Hodge E.J. (2004). Palaeoclimate of the Western Mediterranean region: results from

speleothems. Unpublished Ph.D. Thesis, University of Bristol, 246 p.

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: 713-718

Ivanovich, M., Harmon, R.S. (1992). Uranium-series disequilibrium: applications to earth,

marine, and environmental sciences (2nd edition). United Kingdom: Clarendon Press, 910 p.

Jaffey, A.H., Flynn, K.F., Glendenin, L.E., Bentley, W.C., Essling, A.M. (1971). Precision

measurement of half-lives and specific activities of 235U and 238U. Physical Review C, 4: 1889-

1906.

Just, J., Hübscher, C., Betzler, C., Lüdmann, T., & Reicherter, K. (2011). Erosion of continental

margins in the Western Mediterranean due to sea-level stagnancy during the Messinian Salinity

Crisis. Geo-Marine Letters, 31 (1): 51-64.

Kaufman, A. (1993). An evaluation of several methods for determining 230Th/U ages in impure

carbonates. Geochimica et Cosmochimica Acta, 57: 2303-2317.

Kelly, S.D., Newville, M.G., Cheng, L., Kemner, K.M., Sutton, S.R., Fenter, P., Sturchio, N. C.,

Spötl, C. (2003). Uranyl incorporation in natural calcite, Environmental Science Technology, 37:

1284-1287.

Kemp, A.C., Dutton, A., Raymo, M.E. (2015). Paleo constraints on future sea-level rise. Current

Climate Change Reports, 1 (3): 205-215.

89

Kopp, R.E., Simons, F.J., Mitrovica, J.X., Maloof, A.C., Oppenheimer, M. (2009). Probabilistic

assessment of sea level during the last interglacial stage. Nature, 462: 863-867.

Koster, R.D., De Valpine, D.P., Jouzel, J. (1993). Continental water recycling and H218O

concentrations. Geophysical Research Letters, 20 (20): 2215-2218.

Lachniet, M.S. (2009). Climatic and environmental controls on speleothem oxygen-isotope

values. Quaternary Science Reviews, 28 (5-6): 412-432.

Lambeck, K., Chappell, J. (2001). Sea level change through the last glacial cycle. Science, 292:

679-686.

Langmuir, D. (1978). Uranium solution‐mineral equilibria at low temperatures with applications

to sedimentary ore deposits, Geochimica et Cosmochimica Acta, 42: 547-569.

Langmuir, D., Herman, J.S. (1980). The mobility of thorium in natural waters at low

temperatures. Geochimica et Cosmochimica Acta, 44: 1753-1766.

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: 534-536.

Lisiecki, L. E., Raymo, M. E. (2005). A Plio-Pleistocene stack of 57 globally distributed benthic

δ18O records. Paleoceanography, 20: 1-17.

Lopez, J.M. (2007). Las manifestaciones hidrotermales del sur de Llucmajor, Mallorca.

Unpublished Memoria d’Investigacio. Palma de Mallorca: Universitat de les Illes Balears,

Departament de Ciencies de la Terra, 132 p.

Lorscheid, T., Stocchi, P., Casella, E., Gómez-Pujol, L., Vacchi, M., Mann, T., Rovere, A.

(2017). Paleo sea-level changes and relative sea-level indicators: Precise measurements,

indicative meaning and glacial isostatic adjustment perspectives from Mallorca (Western

Mediterranean). Palaeogeography, Palaeoclimatology, Palaeoecology, 473: 94-107.

Ludwig, K.R., Simmons, K.R., Szabo, B.J., Winograd, I.J., Landwehr, J.M., Riggs, A.C,

Hoffman, R.J. (1992). Mass-spectrometric 230Th-234U-238U dating of the Devils Hole calcite vein.

Science, 258: 284-287.

Ludwig, K.R. (1998). On the treatment of concordant uranium–lead ages. Geochimica et

Cosmochimica Acta, 62: 665-676.

Ludwig, K.R. (2001). User manual for Isoplot/Ex version 2.05. Berkeley Geochronology Center,

Special Publication, 1a, 2455 p.

Luo, X., Rehkämper, M., Lee, D.C., Halliday, A.N. (1997). High precision 230Th/232Th and 234U/238U measurements using energy filtered ICP magnetic sector multiple collector mass

spectrometry. International Journal of Mass Spectrometry and Ion Processes, 171 (1-3): 105-117.

90

Mas, G., Maillard, A., Alcover, J.A., Fornós, J.J., Bover, P., Torres-Roig, E. (2018). Terrestrial

colonization of the Balearic Islands: New evidence for the Mediterranean sea-level drawdown

during the Messinian Salinity Crisis. Geology, 46 (6): 527-530.

McDermott, F. (2004). Palaeo-climate reconstruction from stable isotope variations in

speleothems: a review. Quaternary Science Reviews, 23 (7-8): 901-918.

McDermott, F., Schwarcz, H.P., Rowe, P.J. (2006). Isotopes in speleothems In: M.J., Leng (ed.),

Isotopes in Palaeoenvironmental Research, Dordrecht: Springer, 185-226.

Mickler, P.J., Banner, J.L., Stern, L., Asmerom, Y., Edwards, R. L., Ito, E. (2004). Stable isotope

variations in modern tropical speleothems: evaluating equilibrium vs. kinetic isotope effects.

Geochimica et Cosmochimica Acta, 68 (21): 4381-4393.

Moorbath, S., Taylor, P.N., Orpen, J.L., Treloar, P., Wilson, J.F. (1987). First direct radiometric

dating of Archaen stromatolitic limestone. Nature, 326: 865-867.

Muhs, D.R., Simmons, K.R., Steinke, B. (2002). Timing and warmth of the Last Interglacial

period: new U-series evidence from Hawaii and Bermuda and a new fossil compilation for North

America. Quaternary Science Reviews, 21 (12-13): 1355-1383.

Mylroie, J.E., Mylroie, J.R. (2007). Developments of the Carbonate Island Karst Model. Journal

of Cave and Karst Studies, 69: 59-75.

Naish, T., Powell, R., Levy, R., Wilson, G., Scherer, R., Talarico, F., et al. (2009). Obliquity-

paced Pliocene West Antarctic ice sheet oscillations. Nature, 458: 322-328.

Naish, T.R., Wilson, G.S. (2008). Constraints on the amplitude of Mid-Pliocene (3.6-2.4 Ma)

eustatic sea-level fluctuations from the New Zealand shallow-marine sediment record.

Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering

Sciences, 367: 169-187.

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’Historia Natural de les Balears, 18: 147-162.

Palmer, A.N. (1991). Origin and morphology of limestone caves. Geological Society of America

Bulletin, 103 (1): 1-21.

Pickering, R., Kramers, J.D., Partridge, T., Kodolanyi, J., Pettke, T. (2010). U–Pb dating of

calcite–aragonite layers in speleothems from hominin sites in South Africa by MC-ICP-MS.

Quaternary Geochronology, 5 (5): 544-558.

Pickering, R., Dirks, P.H.G.M., Jinnah, Z., de Ruiter, D.J., Churchill, S.E., Herrires, A.I.R.,

Woodhead, J.D., Hellstrom, J.C., Berger, L.R. (2011). Australopithecus sediba at 1.977 Ma and

implications for the origins of genus Homo. Science, 333: 1421-1423.

91

Pickering, R., Kramers, J.D., Hancox, P.J., de Ruiter, D.J., Woodhead, J.D. (2011).

Contemporary flowstone development links early hominin bearing cave deposits in South Africa.

Earth and Planetary Science Letters, 306: 23-32.

Pickering, R., Herries, A.I., Woodhead, J.D., Hellstrom, J.C., Green, H.E., Paul, B., Ritzman, T.,

Strait, D.S., Schoville, B.J., Hancox, P.J. (2019). U–Pb-dated flowstones restrict South African

early hominin record to dry climate phases. Nature, 565: 226-229.

Pollard, D., DeConto, R. M. (2009). Modelling West Antarctic ice sheet growth and collapse

through the past five million years. Nature, 458: 329-332.

Polyak, V.J., Hill, C., Asmerom, Y. (2008). Age and evolution of the Grand Canyon revealed by

U\Pb dating of water-table-type speleothems. Science, 319: 1377-1380.

Polyak, V.J., Asmerom, Y. (2009). Overcoming initial 230Th problems in Uranium-series dating

without using 234U/232Th - 230Th/232Th isochrons. Proceedings of the 15th International Congress

of Speleology, Kerrville, TX, 2 (2): 1060-1064.

Polyak, V.J., DuChene, H.R., Davis, D.G., Palmer, A.N., Palmer, M.V., Asmerom, Y. (2013).

Incision history of Glenwood Canyon, Colorado, USA, from the uranium-series analyses of

water-table speleothems. International Journal of Speleology, 42 (3): 193-202.

Polyak, V.J., Provencio, P.P., Asmerom, Y., (2016). U-Pb dating of speleogenetic dolomite: A new

sulfuric acid speleogenesis chronometer. International Journal of Speleology, 45 (2): 103-109.

Polyak, V.J., Onac, B.P., Fornós, J.J., Hay, C., Asmerom, 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: 860-864.

Pomar, L., Ginés, A., Ginés, J. (1979). Morfología, estructura y origen de los espeleotemas

epiacuáticos. Endins, 5: 3-17.

Pomar, L. (1991). Reef geometries, erosion surfaces and high frequency sea level changes,

Upper Miocene Reef Complex, Mallorca, Spain. Sedimentology, 38: 243-269.

Pomar, L., Ward, W.C., Green, D. (1996). Upper Miocene reef complex of the Llucmajor area,

Mallorca, Spain. In: Franseen, E., Esteban, M., Ward, W.C., Rouch, J.M. (eds.) Models for

carbonate stratigraphy from Miocene reef complexes of Mediterranean regions. SEPM Concepts

in Sedimentology and Paleontology, 5: 191-225.

Ramis, C., Romero, R., Homar, V. (2009). The severe thunderstorm of 4 October 2007 in

Mallorca: an observational study. Natural Hazards and Earth System Sciences, 9 (4): 1237-1245.

Rasmussen, S.O., Bigler, M., Blockley, S.P., Blunier, T., Buchardt, S.L., Clausen, H.B.,

Cvijanovic, I., Dahl-Jensen, D., Johnsen, S.J., Fischer, H., Gkinis, V. et al. (2014). A

stratigraphic framework for abrupt climatic changes during the Last Glacial period based on

three synchronized Greenland ice-core records: refining and extending the INTIMATE event

stratigraphy. Quaternary Science Reviews, 106: 14-28.

92

Raymo, M.E., Lisiecki, L.E., Nisancioglu, K.H. (2006). Plio-Pleistocene ice volume, Antarctic

climate, and the global δ18O record. Science, 313: 492-495.

Reeder, R.J., Lamble, G.M., Northrup, P.A. (1999). XAFS study of the coordination and local

relaxation around Co2+, Zn2+, Pb2+, and Ba2+ trace elements, American Mineralogist, 84: 1049-1060.

Richards, D.A., Smart, P.L., Edwards, R.L. (1994). Maximum sea levels for the last glacial

period from U‐series ages of submerged speleothems. Nature, 367: 357–360.

Richards, D.A., Bottrell, S.H., Cliff, R.A., Strohle, K., Rowe, P.J. (1998). U/Pb dating of a

speleothem of Quaternary age. Geochimica et Cosmochimica Acta, 62: 3683-3688.

Richards, D.A., Dorale, J.A. (2003). Uranium-series chronology and environmental applications

of speleothems. Reviews in Mineralogy and Geochemistry, 52 (1): 407-460.

Robledo P.A., Pomar L. (2000). Upper Miocene karst collapse structures of the east coast,

Mallorca, Spain. Acta Carsologica, 29: 2-12.

Robledo Ardila, P.A., Durán, J.J., Pomar, L. (2004). Paleocollapse structures as geological

record for reconstruction of past karst processes during the Upper Miocene of Mallorca Island.

International Journal of Speleology, 33 (1-4): 81-95.

Rodríguez-Perea, A., Gelabert, B. (1998). Geologia de Mallorca. In: Fornós, J.J. (ed.) Aspectes

Geològics de les Balears. Palma de Mallorca, Universitat de les Illes Balears, pp. 11-38.

Rohling, E.J., Bigg, G.R. (1998). Paleosalinity and δ18O: a critical assessment. Journal of

Geophysical Research: Oceans, 103 (C1): 1307-1318.

Rohling, E.J., Grant, K., Bolshaw, M., Roberts, A.P., Siddall, M., Hemleben, C., Kucera, M.

(2009). Antarctic temperature and global sea level closely coupled over the past five glacial

cycles. Nature Geoscience, 2 (7): 500-504.

Rohling, E.J., Braun, K., Grant, K., Kucera, M., Roberts, A.P., Siddall, M., Trommer, G. (2010).

Comparison between Holocene and Marine Isotope Stage-11 sea-level histories. Earth and

Planetary Science Letters, 291 (1-4): 97-105.

Rohling, E.J., Foster, G.L., Grant, K.M., Marino, G., Roberts, A.P., Tamisiea, M.E., Williams, F.

(2014). Sea-level and deep-sea-temperature variability over the past 5.3 million years. Nature,

508: 477-482.

Rosholt, J.N., Antal, P.S. (1962). Evaluation of the 231Pa/U-230Th/U method for dating

Pleistocene carbonate rocks. US Geological Survey Professional Paper, 450: 108-111.

Rosselló, V.M., Balaguer, P., Fornós, J.J., Ginés, J., Gómez-Pujol, Ll., Pardo-Pascual, J.E.,

Segura, F. (2007). El modelado de la costa rocosa oriental de Mallorca: macro, meso y

microformas. In: Fornós, J.J., Ginés, J., Gómez-Pujol, Ll. (eds.). Geomorfología Litoral:

Llevant y Migjorn de Mallorca. Monografies de la Societat. d’Historia Natural de les Balears,

15: 153-172.

93

Rovere, A., Raymo, M.E., Vacchi, M., Lorscheid, T., Stocchi, P., Gomez-Pujol, L., Harris, D.L.,

Casella, E., O’Leary, M.J., Hearty, P.J. (2016). The analysis of Last Interglacial (MIS 5e)

relative sea-level indicators: Reconstructing sea-level in a warmer world. Earth-Science Reviews,

159: 404-427.

Rovere, A., Stocchi, P., Vacchi, M. (2016). Eustatic and relative sea level changes. Current

Climate Change Reports, 2 (4): 221-231.

Rudzka, D., McDermott, F., Baldini, L.M., Fleitmann, D., Moreno, A., Stoll, H. (2011). The

coupled δ13C-radiocarbon systematics of three Late Glacial/early Holocene speleothems; insights

into soil and cave processes at climatic transitions. Geochimica et Cosmochimica Acta, 75 (15):

4321-4339.

Rutherford, E., Soddy, F. (1902). The radioactivity of thorium compounds II. The cause and

nature of radioactivity. Journal of the Chemical Society, Transactions, 81: 837-860.

Salomons, W., Mook, W.G. (1986). Isotope geochemistry of carbonates in the weathering zone.

In: Fontes, J.C., Fritz, P. (eds), Handbook of environmental isotope geochemistry: the terrestrial

environment. Amsterdam: Elsevier, 8: 239-270.

Schoene, B. (2014). 4.10 U–Th–Pb Geochronology. In: Turekian, H.D.H.K. (ed.) Treatise on

Geochemistry (2nd edition). Oxford: Elsevier, pp. 341-378.

Schwarcz, H.P. (1986). Geochronology and isotopic geochemistry of speleothems. In: Fontes,

J.C., Fritz, P. (eds), Handbook of environmental isotope geochemistry: the terrestrial

environment. Amsterdam: Elsevier, 8: 271-303.

Shackleton, N.J. (1987). Oxygen isotopes, ice volume and sea level. Quaternary Science

Reviews, 6 (3-4): 183-190.

Shen, C.C., Edwards, R.L., Cheng, H., Dorale, J.A., Thomas, R.B., Moran, S.B., Weinstein, S.E,

Edmonds, H.N. (2002). Uranium and thorium isotopic and concentration measurements by

magnetic sector inductively coupled plasma mass spectrometry. Chemical Geology, 185 (3-4):

165-178.

Siddall, M., Rohling, E.J., Almogi-Labin, A., Hemleben, C., Meischner, D., Schmelzer, I.,

Smeed, D.A. (2003). Sea-level fluctuations during the last glacial cycle. Nature, 423: 853-858.

Smart, P.L., Friedrich, H. (1987). Water movement and storage in the unsaturated zone of a

maturaly karstified aquifer, Mendip Hills, England. Proceedings of the Conference on

Environmental Problems in Karst Terrains and their Solution, Bowling Green, Kentucky.

National Water Well Association, pp. 57-87.

Spratt, R.M., Lisiecki, L.E. (2016). A Late Pleistocene sea level stack. Climate of the Past, 12

(4): 1079-1092.

Steiger, R.H., Jäger, E. (1977) Subcommission on Geochronology: Convention on the use of decay

constants in geo- and cosmochronology. Earth and Planetary Science Letters, 36: 359-362.

94

Stocchi, P., Vacchi, M., Lorscheid, T., de Boer, B., Simms, A.R., van de Wal, R.S., Vermeersen,

B.L.A., Pappalardo, M., Rovere, A. (2018). MIS 5e relative sea-level changes in the

Mediterranean Sea: Contribution of isostatic disequilibrium. Quaternary Science Reviews, 185:

122-134.

Sturchio, N.C., Antonio, M.R., Saderholm, L., Sutton, S.R., Brannon, J.C. (1998). Tetravalent

uranium in calcite. Science, 281: 971-973.

Suess, E. (1909). The face of the earth (Das antlitz der erde) (Vol. 4). Clarendon Press.

Sumner, G., Homar, V., Ramis, C. (2001). Precipitation seasonality in the Eastern and Southern

Coastal Spain. International Journal of Climatology, 21: 219-247.

Tera, F., Wasserburg, G.J. (1972). U–Th–Pb systematics in three Apollo 14 basalts and the

problem of initial Pb in lunar rocks. Earth and Planetary Science Letters, 4: 281-304.

Thomson, J.J. (1914). Rays of positive electricity, Proceedings of the Royal Society of London,

89: 1.

Thompson, P., Schwarcz, H.P., Ford, D.C. (1974). Continental Pleistocene climatic variations

from speleothem age and isotopic data. Science, 184: 893-895.

Thompson, P., Ford, D.C., Schwarcz, H.P. (1975). 234U/238U ratios in limestone cave seepage

waters and speleothem from West Virginia. Geochimica et Cosmochimica Acta, 39 (5): 661-669.

Tuccimei, P., Antonioli, F., Delitala, M.C., Fornós, J.J., Ginés, A., Ginés, J., Mucedda, M., Orrù,

P., Taddeucci, A. (2000a). Sea level changes during oxygen isotope stage 5 and 7 at capo Caccia

(Alghero, Sardinia), based on U/Th datings of phreatic overgrowths on speleothems and

geomorphological evidences. Proceedings of AIQUA Workshop: ‘‘Tyrrhenian in Italy’’, pp. 24-26.

Tuccimei, P., Ginés, J., Delitala, C., Pazzelli, L., Taddeucci, A., Clamor, B., Fornós, J.J., Ginés,

A., Gràcia, F. (2000b). Dataciones Th/U de espeleotemas freáticos recolectados a cotas inferiores

al actual nivel marino en cuevas costeras de Mallorca (España): Aportaciones a la construcción

de una curva eustática detallada de los últimos 300 ka para el Mediterráneo occidental. Endins,

23: 59-72.

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 (1): 1-

21.

Tuccimei, P.; Soligo, M.; Ginés, J.; Ginés, A.; Fornós, J.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

(7): 782-790.

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. New

York: John Wiley & Sons, pp. 83-103.

95

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

(9): 865-879.

Villa, I.M., Bonardi, M.L., De Bièvre, P., Holden, N.E., Renne, P.R. (2016). IUPAC-IUGS status

report on the half-lives of 238U, 235U and 234U. Geochimica et Cosmochimica Acta, 172: 387-392.

Waelbroeck, C., Labeyrie, L., Michel, E., Duplessy, J. C., McManus, J. F., Lambeck, K., Balbon,

E., Labracherie, M. (2002). Sea-level and deep water temperature changes derived from benthic

foraminifera isotopic records. Quaternary Science Reviews, 21 (1-3): 295-305.

Walker, J., Cliff, R.A., Latham, A.G. (2006). U\Pb isotopic age of the StW 573 hominid from

Sterkfontein, South Africa. Science, 314: 1592-1594.

Wehmiller, J.F., Simmons, K.R., Cheng, H., Edwards, R.L., Martin-McNaughton, J., York, L.L.,

Krantz, D.E., Shen, C.C. (2004). Uranium-series coral ages from the US Atlantic Coastal Plain–

the “80 ka problem” revisited. Quaternary International, 120 (1): 3-14.

Wendt, I. (1984). A three-dimensional U/Pb discordia plane to evaluate samples with common

lead of unknown isotopic composition. Chemical Geology, 46: 1-12.

Wetherill, G.W. (1956). Discordant uranium-lead ages. Transactions of American Geophysical

Union, 37: 320-326.

White, W.M. (2015). Isotope Geochemistry (1st edition). Oxford: John Wiley & Sons, 496 p.

Woodhead, J., Hellstrom, J., Maas, R., Drysdale, R., Zanchetta, G., Devine, P., Taylor, E. (2006).

U-Pb geochronology of speleothems by MC-ICPMS. Quaternary Geochronology, 1 (3): 208-221.

Woodhead, J., Pickering, R. (2012). Beyond 500 ka: progress and prospects in the U-Pb

chronology of speleothems, and their application to studies in palaeoclimate, human evolution,

biodiversity and tectonics. Chemical Geology, 322-323: 290-299.

Woodhead, J.D., Hellstrom, J., Pickering, R., Drysdale, R., Paul, B, Bajo, P. (2012). U and Pb

variability in older speleothems and strategies for their chronology. Quaternary Geochronology,

14: 105-113.

Zazo, C., Goy, J.L., Dabrio, C. J., Bardajı, T., Hillaire-Marcel, C., Ghaleb, B., González-

Delgado, J.Á, Soler, V. (2003). Pleistocene raised marine terraces of the Spanish Mediterranean

and Atlantic coasts: records of coastal uplift, sea-level highstands and climate changes. Marine

Geology, 194 (1-2): 103-133.

Zazo, C., Goy, J.L., Dabrio, C.J., Lario, J., González-Delgado, J.A., Bardají, T., Hillaire-Marcel,

C., Cabero, A., Ghaleb, B., Borja, F., Silva, P.G. (2013). Retracing the Quaternary history of sea-

level changes in the Spanish Mediterranean–Atlantic coasts: Geomorphological and

sedimentological approach. Geomorphology, 196: 36-49.

Zheng, Y. F. (1992) The three-dimensional U/Pb method: Generalized models and implications

for U–Pb two-stage systematics. Chemical Geology, 100: 3-18.

96

APPENDIX A:

DATA

97

Table A1. U-series dataset and ages of the samples in this study*.

* Initial 230Th/232Th corrections are: a Computed from Eq. 12; b Assumed 230Th/232Th atomic ratio of 2.0 × 10-5 ppm ±50%; c Assumed 230Th/232Th

atomic ratio of 4.4 × 10-5 ppm ± 50%. Not marked samples are ages corrected using an initial 230Th/232Th atomic ratio of 4.4 ppm ± 50%. Ages are

reported in years before present (BP), where the present is 2018. All errors are absolute 2σ.

Location Sample CodeElevation

(mapsl)

Corrected

Age

(yr BP)

errors

(+)errors (-)

CF-10-2 2 m 388 ± 0.5 1259 ± 51 956 ± 39 1.0164 ± 0.0025 1.022 ± 0.0016 22 ± 2 83 ± 7 465132 +26402 -22491CF-12-1 2.9 m 353 ± 0.5 6934 ± 47 158 ± 1 1.0121 ± 0.0025 1.013 ± 0.0017 13 ± 2 62 ± 13 558020 +75337 -50357CF-13-1 2.8 m 331 ± 0.4 20 ± 30 51183 ± 75235 1.0231 ± 0.0024 1.020 ± 0.0020 18 ± 1 653948 +31672 -30346CF-14-1 3 m 256 ± 0.4 32 ± 57 24439 ± 43175 1.0135 ± 0.0032 1.017 ± 0.0022 17 ± 2 70 ± 12 500322 +51562 -38491

CF-11-35 a 298 ± 0.5 72 ± 10 8156 ± 1134 0.6461 ± 0.0020 1.167 ± 0.0026 167 ± 3 212 ± 3 85690 +512 -509

CF-11-21 a 229 ± 0.3 7368 ± 10 61 ± 0.2 0.6445 ± 0.0018 1.152 ± 0.0023 152 ± 2 192 ± 3 84097 +1673 -1649

CF-11-18 a 393 ± 0.5 10177 ± 9 77 ± 0.2 0.6530 ± 0.0017 1.163 ± 0.0017 163 ± 2 207 ± 2 85452 +1190 -1178

CF-11-13 a 225 ± 0.3 2822 ± 4 155 ± 0.5 0.6388 ± 0.0019 1.150 ± 0.0020 150 ± 2 191 ± 3 84666 +950 -942

CF-11-10 a 283 ± 0.4 281 ± 4 1962 ± 31 0.6382 ± 0.0015 1.163 ± 0.0018 163 ± 2 207 ± 2 84349 +405 -403

DIP-1-A-1 393 ± 0.5 2664 ± 49 460 ± 8 1.0189 ± 0.0026 1.009 ± 0.0015 9.3 ± 1 646459 +55877 -51727DIP-1-A-2 313 ± 0.4 1096 ± 2 890 ± 2 1.0180 ± 0.0024 1.009 ± 0.0012 8.7 ± 1 602644 +62150 -57393DIP-1-B-1 3 m 458 ± 0.6 17 ± 1 84347 ± 4139 1.0126 ± 0.0024 1.007 ± 0.0011 6.6 ± 1 613333 +81615 -73455FE-1-F-2 5.5 m 202 ± 0.2 4597 ± 32 135 ± 1 1.0033 ± 0.0024 1.011 ± 0.0015 11 ± 2 44 ± 7 478301 +29487 -24658FE-1-E-1 4.5 m 385 ± 0.5 168 ± 2 6989 ± 94 0.9991 ± 0.0024 0.999 ± 0.0014 -1 ± 1 583942 +88305 -84743MIP-1-1 7 m 80 ± 0.2 654 ± 47 371 ± 27 0.9982 ± 0.0038 1.010 ± 0.0034 10 ± 3 37 ± 13 450976 +46187 -34984MIP-2-1 7 m 139 ± 0.3 3037 ± 4 140 ± 0.4 1.0026 ± 0.0030 1.008 ± 0.0028 8 ± 3 35 ± 14 515426 +74049 -49002MIP-3-1 10.8 m 82 ± 0.1 5727 ± 52 44 ± 0.4 1.0076 ± 0.0040 1.007 ± 0.0029 7 ± 3 579998 +76635 -68888MIP-4A-1 10.2 m 176 ± 0.3 1085 ± 61 497 ± 28 1.0026 ± 0.0032 1.003 ± 0.0026 3 ± 3 590577 +78492 -70262MIP-4B-1 10.4 m 115 ± 0.2 203 ± 49 1731 ± 415 0.9996 ± 0.0040 1.007 ± 0.0032 7 ± 3 27 ± 16 507683 +88854 -55352MIP-5-25 35 ± 0.1 5015 ± 40 22 ± 0.2 1.0386 ± 0.0065 1.013 ± 0.0033 13 ± 3 1080999 +63139 -61036

MIP-5-18 b 179 ± 0.2 13565 ± 28 43 ± 0.1 1.0693 ± 0.0024 1.106 ± 0.0017 106 ± 2 249 ± 5 296210 +2733 -2654

MIP-5-16 b 227 ± 0.3 3435 ± 45 218 ± 2.9 1.0783 ± 0.0027 1.125 ± 0.0020 125 ± 2 280 ± 6 284449 +2818 -2733

MIP-5-8 b 201 ± 0.3 16439 ± 46 39 ± 0.1 1.0363 ± 0.0026 1.139 ± 0.0020 139 ± 2 261 ± 4 223532 +1509 -1481

MIP-5-3 b 271 ± 0.4 3832 ± 35 216 ± 2.0 1.0001 ± 0.0024 1.179 ± 0.0016 179 ± 2 303 ± 3 185904 +1542 -1524

DR-D20-3A c 3266 ± 4.3 8373 ± 11 554 ± 1 0.4650 ± 0.0011 1.160 ± 0.0012 160 ± 1 187 ± 1 54466 +75 -75

DR-D20-3B c 4815 ± 5.6 31605 ± 13 247 ± 1 0.5306 ± 0.0012 1.160 ± 0.0012 160 ± 1 191 ± 1 63968 +92 -92

DR-D20-3C c 2499 ± 4.5 4053 ± 106 948 ± 25 0.5029 ± 0.0014 1.159 ± 0.0014 159 ± 1 189 ± 2 60695 +323 -322

DR-D20-6 c 3129 ± 2.8 4711 ± 72 1288 ± 20 0.6344 ± 0.0014 1.163 ± 0.0012 163 ± 1 206 ± 1 83663 +354 -353

DR-D20-4S 135 ± 0.2 1113 ± 57 359 ± 57 0.9655 ± 0.0027 1.122 ± 0.0025 122 ± 3 214 ± 5 198923 +2116 -2082

DR-D20-16.5 mm c 328 ± 0.3 755 ± 2 1296 ± 4 0.9771 ± 0.0022 1.112 ± 0.0011 112 ± 1 204 ± 2 211817 +755 -747

DR-D20-12.5 mm c 371 ± 0.4 36 ± 1 30839 ± 1022 0.9903 ± 0.0022 1.118 ± 0.0011 118 ± 1 217 ± 2 216704 +793 -785

DR-D20-3 mm c 286 ± 0.5 105 ± 2 8258 ± 196 0.9928 ± 0.0027 1.114 ± 0.0014 114 ± 1 213 ± 3 221155 +1043 -1029

DR-D20-straw c 228 ± 0.2 252 ± 3 2757 ± 31 0.9954 ± 0.0023 1.111 ± 0.0014 111 ± 1 210 ± 3 224736 +1092 -1077

1 m

Cala Falcó

model age =

model age =

model age =

model age =

Dimoni

238U

(ppb)

232Th

(ppt)

Initial

δ234

U

(‰)

234U/

238U

activity

230Th/

232Th

activity

230Th/

238U

activity

Measured

δ234

U

(‰)

sa

Ferradura

5 m

model age =

na Mitjana

12.3 m

-1.1 mCoves del

Drac

model age =

model age =

model age =

98

Table A2. U-Pb data for the flowstone FE-1-A.

Location Sample CodeElevation

(mapsl)

FE-1-A-1-Pb-A 74 ± 0.1 91 ± 34 27 ± 0.10 18 ± 0.102 7.6 ± 0.205 139 ± 0.2 1.003 ± 0.008 1.003 ± 0.002 3.0 ± 1.5

FE-1-A-1-Pb-B 95 ± 0.1 79 ± 41 276 ± 0.05 21 ± 0.101 0.9 ± 0.174 18 ± 0.1 0.994 ± 0.005 1.003 ± 0.001 3.2 ± 1.4

FE-1-A-1-Pb-C 46 ± 0.1 33 ± 30 541 ± 0.14 18 ± 0.102 1.3 ± 0.233 23 ± 0.2 1.023 ± 0.007 1.012 ± 0.002 12.2 ± 2.1

FE-1-A-1-Pb-D 149 ± 0.2 121 ± 31 208 ± 0.06 18 ± 0.101 2.0 ± 0.186 36 ± 0.2 0.988 ± 0.005 1.004 ± 0.001 3.8 ± 1.3

FE-1-A-1-Pb-E 74 ± 0.1 97 ± 40 49 ± 0.06 18 ± 0.102 4.2 ± 0.216 76 ± 0.2 0.993 ± 0.006 1.005 ± 0.002 4.8 ± 2.5

FE-1-A-2-Pb-A 11 ± 0.1 102 ± 44 153 ± 0.06 18 ± 0.121 1.2 ± 0.243 22 ± 0.2 0.997 ± 0.009 1.001 ± 0.002 1.3 ± 2.4

FE-1-A-2-Pb-B 84 ± 0.1 191 ± 39 79 ± 0.05 18 ± 0.103 2.9 ± 0.198 53 ± 0.2 0.993 ± 0.008 1.003 ± 0.002 3.3 ± 1.9

FE-1-A-2-Pb-C 73 ± 0.1 44 ± 27 104 ± 0.14 18 ± 0.121 1.9 ± 0.230 35 ± 0.2 0.996 ± 0.005 1.005 ± 0.001 5.2 ± 1.4

FE-1-A-2-Pb-D 84 ± 0.1 72 ± 33 199 ± 0.06 18 ± 0.101 1.1 ± 0.186 20 ± 0.2 0.991 ± 0.007 1.001 ± 0.002 0.7 ± 1.6

FE-1-A-2-Pb-E 95 ± 0.1 179 ± 33 9 ± 0.16 19 ± 0.129 29.7 ± 0.310 552 ± 0.3 0.996 ± 0.008 1.006 ± 0.002 5.7 ± 2.3

3.5 m

206Pb/

204Pb

Total Pb

(ppb)

232Th

(ppt)

238U

(ppb)238

U/206

Pb 238

U/204

Pb 230

Th/238

U

activity

234U/

238U

activity

Measured

δ234

U

(‰)

sa

Ferradura

99

Table A3. Carbon and Oxygen isotopic data for CF-11 stalagmite.

Continue next page.

Depth from top

(mm)

Age

(ka)

δ13

C

(‰ VPDB)

δ18

O

(‰ VPDB)

0.0 83.69 -5.23 -4.06

0.5 83.72 -1.03 -3.35

1.0 83.75 -5.53 -3.28

1.5 83.79 -3.00 -2.80

2.0 83.82 -5.20 -2.78

2.5 83.86 -4.86 -3.25

3.0 83.89 -4.36 -3.08

3.5 83.92 -3.11 -3.12

4.0 83.96 -4.71 -2.58

4.5 83.99 -5.07 -2.46

5.0 84.02 -5.99 -2.95

5.5 84.06 -4.41 -2.35

6.0 84.09 -5.11 -2.96

6.5 84.13 -7.07 -3.19

7.0 84.16 -5.57 -2.92

7.5 84.19 -6.61 -3.31

8.0 84.23 -5.65 -2.58

8.5 84.26 -5.27 -2.65

9.0 84.30 -5.07 -2.66

9.5 84.34 -4.85 -2.97

10.0 84.38 -6.05 -2.61

10.5 84.43 -2.33 -3.18

11.0 84.47 -4.11 -3.52

11.5 84.52 -2.17 -2.46

12.0 84.57 -1.29 -2.49

12.5 84.61 -2.84 -3.10

13.0 84.64 -4.62 -3.06

13.5 84.68 -4.61 -3.05

14.0 84.71 -1.83 -2.00

14.5 84.74 -3.23 -2.27

15.0 84.76 -6.10 -3.50

15.5 84.79 -3.96 -2.86

16.0 84.82 -2.62 -3.83

100

Depth from top

(mm)

Age

(ka)

δ13

C

(‰ VPDB)

δ18

O

(‰ VPDB)

16.5 84.84 -6.31 -3.94

17.0 84.87 -5.43 -3.76

17.5 84.90 -6.35 -3.08

18.0 84.92 -5.28 -2.75

18.5 84.95 -4.74 -3.47

19.0 84.97 -4.15 -3.45

19.5 84.99 -2.79 -2.92

20.0 85.00 -3.22 -2.69

20.5 85.02 -3.17 -2.96

21.0 85.03 -3.94 -3.18

21.5 85.05 -4.38 -4.52

22.0 85.06 -4.93 -3.48

22.5 85.08 -6.07 -3.87

23.0 85.10 -3.68 -3.41

23.5 85.12 -5.62 -2.93

24.0 85.14 -5.47 -3.35

24.5 85.17 -3.95 -3.91

25.0 85.19 -5.13 -3.17

26.0 85.24 -5.88 -3.65

26.5 85.26 -5.49 -2.83

27.0 85.29 -6.27 -3.47

27.5 85.31 -5.25 -2.79

28.0 85.33 -5.94 -3.27

28.5 85.36 -6.24 -4.14

29.0 85.38 -6.86 -3.84

29.5 85.41 -6.53 -3.75

30.0 85.43 -5.38 -3.18

30.5 85.46 -5.36 -3.23

31.0 85.49 -5.75 -3.78

31.5 85.51 -5.29 -3.23

32.0 85.53 -5.33 -2.77

32.5 85.56 -4.89 -2.50

33.0 85.58 -4.97 -2.72

33.5 85.60 -6.25 -3.83

34.0 85.63 -6.67 -3.22

34.5 85.65 -6.55 -5.01

35.0 85.68 -6.42 -3.94

35.5 85.70 -5.95 -3.97

36.0 85.73 -5.51 -3.07

101

Table A4. Age-depth model dataset for CF-11 stalagmite generated from StalAge algorithm

(Scholz and Hoffmann, 2011).

Continue next page.

Depth from top

(mm)

Age

(ka)

Error (+)

(ka)

Error (-)

(ka)

0 83.69 84.55 81.76

0.5 83.72 84.55 81.88

1 83.75 84.56 81.99

1.5 83.79 84.57 82.11

2 83.82 84.57 82.23

2.5 83.86 84.58 82.34

3 83.89 84.59 82.46

3.5 83.92 84.60 82.57

4 83.96 84.60 82.69

4.5 83.99 84.61 82.80

5 84.02 84.62 82.91

5.5 84.06 84.62 83.02

6 84.09 84.64 83.14

6.5 84.13 84.65 83.24

7 84.16 84.67 83.35

7.5 84.19 84.69 83.47

8 84.23 84.70 83.58

8.5 84.26 84.70 83.70

9 84.30 84.70 83.81

9.5 84.34 84.72 83.91

10 84.38 84.77 83.98

10.5 84.43 84.85 84.02

11 84.47 84.96 84.04

11.5 84.52 85.07 84.05

12 84.57 85.17 84.07

12.5 84.61 85.23 84.09

13 84.64 85.27 84.12

13.5 84.68 85.31 84.15

14 84.71 85.34 84.17

14.5 84.74 85.40 84.17

15 84.76 85.46 84.17

15.5 84.79 85.54 84.17

16 84.82 85.61 84.17

102

Depth from top

(mm)

Age

(ka)

Error (+)

(ka)

Error (-)

(ka)

16.5 84.84 85.65 84.18

17 84.87 85.69 84.18

17.5 84.90 85.73 84.20

18 84.92 85.77 84.23

18.5 84.95 85.82 84.25

19 84.97 85.86 84.28

19.5 84.99 85.89 84.32

20 85.00 85.89 84.34

20.5 85.02 85.89 84.36

21 85.03 85.89 84.38

21.5 85.05 85.89 84.41

22 85.06 85.89 84.44

22.5 85.08 85.89 84.48

23 85.10 85.89 84.54

23.5 85.12 85.89 84.59

24 85.14 85.89 84.64

24.5 85.17 85.89 84.68

25 85.19 85.90 84.71

26 85.24 85.92 84.76

26.5 85.26 85.93 84.79

27 85.29 85.94 84.82

27.5 85.31 85.95 84.85

28 85.33 85.96 84.88

28.5 85.36 85.97 84.91

29 85.38 85.99 84.93

29.5 85.41 86.00 84.96

30 85.43 86.01 84.98

30.5 85.46 86.03 85.00

31 85.49 86.04 85.02

31.5 85.51 86.05 85.04

32 85.53 86.07 85.05

32.5 85.56 86.08 85.07

33 85.58 86.10 85.09

33.5 85.60 86.11 85.10

34 85.63 86.13 85.12

34.5 85.65 86.15 85.13

35 85.68 86.18 85.14

35.5 85.70 86.22 85.15

36 85.73 86.26 85.16

103

APPENDIX B:

COPYRIGHT PERMISSIONS

Figure B1. Authorship permission provided by Joaquin Ginés for using a picture from Mondragó

Natural Park in the present study (Fig. 5).

Figure B2. Permission granted from Joaquin Ginés to use his images from Mallorca Island in the present

study (Figs. 12 and 29)

104

Figure B3. Permission provided from Joan Fornós to use some pictures in this thesis (Fig. 29 and 31)

Figure B4. Permission granted by Judy Mcllrath to use some pictures of Mallorca Island in this study

(Figs. 8A, 9B-D, and 29A).

105

Figure B5. Permission granted by Bogdan P. Onac to use the pictures of his property in the present study

(Figs. 4, 8, 9, 10, 11, 28, 29, and 30).

106

Figure B6. Permission obtained by Elsevier to use a picture (Fig. 6 in the present study) originally

published in the book Treatise on Geochemistry.